Skip to main content

Full text of "Mathematics Of Physics And Modern Engineering"

See other formats




library 

Class No. 

Book No. 

Accession No. 




DELHI COLLEGE OF ENGINEERING 

Kashmere Gate, Delhi 

LIBRARY 
DUE DATE 

For each day'e delay after the due date a fine of 50 P. 
per Vol. Shall be charged for books of General Section 
and Be 1.00 for text books Section. 




MATHEMATICS OF PHYSICS 
AND MODERN ENGINEERING 




MATHEMATICS 


OF PHYSICS AND 
MODERN ENGINEERING 

I. S. Sokolnikoff 

Professor of M athematics 
University of California , Los Angeles 

R. M. Redheffer 

Associate Professor of Mathematics 
University of California , Los Angeles 


INTERNATIONAL SIXTH N7 L1HTION 


McGRAW-HILL BOOK COMPANY, INC. 

New York Toronto London 

KOOAKUSHA COMPANY, LTD. 

Tokyo 



MA I'HhMA'1 1C;S OP PHYSICS ANI) 

MODI RN ENGINEERING 

lx n nx i / /ox il \ivm \ / / di nox 

Exclusive lights h\ kdgakusha Co, l.td, jor umnujadun and 
export from Japan I Ins hook i an not ht re exported from 
the country to which it is consigned by A ogaUushn to , ltd 

Copyright © 1958 by t he McGraw-llill Hook Compam, tin 
All lights teseivcd Plus book, oi pails thereof, m.tv not be 
reproduced in any lot in without permission ot the publishers 
Libiary ot Congiess (Catalog Caul Number 57 12914 


'IOSHO INSAISIJ NtlNriM, (.() , LTD. lOKYO, JAPAN 



PREFACE 


The rapidly decreasing time lag between scientific discoveries and appli- 
cations imposes ever-increasing demands on the mathematical equipment 
of scientists and engineers. Although the mathematical preparation of 
engineering students has been strengthened materially in the past thirty 
years, the introduction of courses beyond the tiaditional “terminal course” 
in calculus has been largely confined to a few leading institutions. The 
reluctance to broaden significantly the program of instruction in mathe- 
matics can be attributed in part to the crowded engineering curricula, in 
part to the failure to sense the central position of mathematics in sciences 
and technology, and in part to the scarcity of suitable staffs and instruc- 
tional media. The broadening, however, is inevitable, for it is now gen- 
erally recognized that no professional engineer can keep abreast of scien- 
tific developments without substantially extending his mathematical hori- 
zons. 

This book, in common with its predecessor written by the senior author 
some twenty-five years ago, has as its main aim a sound extension of such 
horizons. The authors not only have been guided by their subjective 
appraisal of the live present-day needs of the engineering profession but 
have also taken into account the views of the leaders of engineering 
thought as expressed in numerous conferences and symposia on engineer- 
ing education sponsored by the National Science Foundation, the American 
Society of Engineering Education, and its predecessor the Society for the 
Promotion of Engineering Education. 

There are many conflicting and often prejudiced currents of thought as 
to how mathematics should be presented to students of applied sciences. 
Some believe that mathematics is one whole and indivisible and hence 
should be presented unto all alike, regardless of the differing creeds. Others 
are content with a catalogue of useful formulas, rules, and devices for 
solving problems. The authors think that these two extremo viewpoints 
are somewhat limited, since they recognize only two of the many facets of 
mathematics. A preoccupation with the logic of mathematics and the over- 
emphasis of a convention called rigor are among the best known means for 
stifling interest in mathematics as a crutch to common sense. On the other 
hand, a presentation which puts applications above the medium making 



VI 


PREFACE 


applications possible is sterile, because it gives no inkling of the supreme 
importance of generalizations and abstractions in applications. The au- 
thors have tried to strike a balance which would make this book both a 
sound and an inspiring introduction to applied mathematics. 

The material in this book appears in nine chapters, each of which is 
complete and virtually independent of the others. Occasional cross refer- 
ences to other chapters are intended to correlate the topics and to enhance 
the usefulness of the book as a reference volume. Each chapter is sub- 
divided into functional parts, many of which also form an organized whole. 
The earlier parts of each chapter are less advanced and should servo as 
an introduction to more difficult topics treated in the later parts. The 
text material set in small type usually deals with generalizations and de- 
velops the less familiar concepts which are sure to grow in importance in 
applications. 

The choice of topics is based on the authors' estimate of the frequency 
with which the subjects treated occur in applications. The illustrative 
material, examples, and problems have been chosen more for their value 
in emphasizing the underlying principles than as a collection of instances 
of dramatic uses of mathematics in specific situations confronting prac- 
ticing engineers. 

Although the book is written so as to require little, if any, outside help, 
the reader is cautioned that no amount of exposition can serve as a substi- 
tute for concentration in following the course of the argument in a serious 
discipline. In order to facilitate the understanding of the principles and 
to cultivate the art of formulating physical problems in the language of 
mathematics, numerous illustrative examples are worked out in detail. 
The authors believe with Newton that exempla nan minus doceunt quam 
precaepta. 

L S. Sokolnikoff 
R. M. Redheffer 



TO THE INSTRUCTOR 


In the sense that a working course in calculus is the sole technical pre- 
requisite, this book is suitable for the beginner in applied mathematics. 
But when viewed in the light of the present-day requirements of the engi- 
neering profession, the text includes a large amount of material of direct 
interest to practicing engineers. 

It is certain that within the next twenty years the methods of functional 
analysis and, in particular, the Hilbert space theory will be in general use 
in technology. A foundation for the assimilation of the function-space 
concepts should be laid now, and we did not hesitate to do so in several 
places in this book. 

We have arranged the contents in nine independent chapters which, in 
turn, are subdivided into parts, most of which can be read independently 
of the rest. The earlier parte of each chapter are less advanced, and our 
experience has shown that several introductory courses for students of sci- 
ence and technology can be based on the material contained in the earlier 
parte. When taken in sequence, this book has ample substance for four 
consecutive semester courses meeting three hours a week. 

This book is also suitable for courses in mathematical analysis bearing 
such labels as ordinary differential equations, partial differential equations, 
vector analysis, advanced calculus, complex variable, and so on. 

Thus Chap. 1, when supplemented by Secs. 12 to 14 of Chap. 2, has 
adequate material for a solid semester course in ordinary differential equa- 
tions. Instructors wishing to include an introduction to numerical meth- 
ods of solutions of differential equations will find suitable material in Secs. 
14 to 18 of Chap. 9. The use of Laplace transforms in solving differential 
equations is discussed in Appendix B, which includes, among other things, 
a meaningful introductory presentation of the “Dirac delta function.’ ' 

Chapter 6, together w r ith Secs. 18 to 25 of Chap. 2, has ample material 
for a semester course in partial differential equations. 

Chapters 4 and 5 have sufficient content for a modem course in vector 
analysis. 

Chapter 7, preceded by the relevant topics on line integrals in Chap. 5, 
is adequate for an introductory course in complex variable theory. 

Chapter 8 can be used in a semester course on probability theory and 



viii 


TO THE INSTRUCTOR 


applications meeting two hours a week. A course entitled “Probability 
and Numerical Methods” meeting three hours a week can be based on 
the material in Chaps. 8 and 9. 

Although this book was written primarily for students of physical sci- 
ences, it is unlikely that a liberal arts student who followed it in an ad- 
vanced calculus course would be obliged to “unlearn" anything in his 
subsequent studies. 

The contents of this book include what we believe should be the mini- 
mum mathematical equipment of a scientific engineer. It may not be out 
of place to note that the mathematical preparation of physicists and engi- 
neers in Russia exceeds the minimum laid down here. While the curricula 
of only a few leading American engineering colleges provide now for mure 
than one year of mathematics beyond calculus, their number will continue 
to increase with the realization that the time allotted to mathematics is a 
sound capital investment, yielding excellent returns both in the time gained 
in professional studies and in the depth of penetration. 



CONTENTS 


Preface v 

To the Instructor vii 

CHAPTER 

1 Ordinary Differential Equations 1 

2 Infinite Series ' >107 

3 Functions of Several Variables 213 

4 Algebra and Geometry of Vectors. Matrices 283 

5 Vector Field Theory 353 

6 Partial Differential Equations 421 

7 Complex Variable 523 

8 Probability 605 

9 Numerical Analysis 673 

APPENDIX 

A Determinants 74 1 

B The Laplace Transform 754 

C Comparison of the Riemann and Lebesgue Integrals 771 

D Table of Hz) - — C / e~‘’ /2 dt 776 

V 2?T J 0 

Answers 777 

Index 799 


lx 




CHAPTER 1 

ORDINARY DIFFERENTIAL EQUATIONS 




Preliminary Remarks and Orientation 

I. Definition of Terms and Generalities 5 

2 Tin* Slipping of a Belt on a Pulley 1 1 

3. Growth 12 

l Diffusion and Chemical Combination 14 

5 The Elastic Curve 15 

The Solution of First-order Equations 

0 Kquat ions w ith Separable Variables 17 

7 Homogeneous Diffeiential Equations 18 

<S Exact Differential Equations 20 

9 Integrating Factors 22 

JO The First-order Linear Equation 23 

1 1. Equations Solvable tor y or y' 25 

12 The Method of Substitution 27 

13. Reduction of Order 29 

Geometry and the First-order Equation 

M Orthogonal Trajectories 30 

15. Parabolic Mirror. Pursuit Curves 33 

16. Singular Solutions 34 

17. The General Behavior of Solutions 36 

Applications of First-order Equations 

18 The Hanging Cham 40 

19. Newton's Law of Motion 42 

20. Newton's Law of Gravitation 46 

Linear Differential Equations 

21. Fan ear Homogeneous Second-order Equations 51 

22 Homogeneous Second-order Linear Equations with Constant 

Coefficients 54 


3 



4 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

23. Differential Operators 57 

24. N onhomogeneous Second-order Linear Equations 59 

25. The Use of Complex Forms of Solutions in Evaluating Particular 

Integrals 64 

26. Linear nth-order Equations with Constant Coefficients 66 

27. General Linear Differential Equations of nth Order 70 

28. Variation of Parameters 72 

29. Reduction of the Order of Linear Equations 76 

30. The Euler-Cauchy Equation 78 

Applications of Linear Equations 

31. Free Vibrations of Electrical and Mechanical Systems 79 

32. Viscous Damping 82 

33. Forced Vibrations. Resonance 86 

34. The Euler Column. Rotating Shaft 90 

Systems of Equations 

35. Reduction of Systems to a Single Equation 95 

36. Systems of Linear Equations with Constant Coefficients 100 



The power and effectiveness of mathematical methods in the study of 
natural sciences stem, to a large extent, from the unambiguous language 
of mathematics, with the aid of which the laws governing natural phe- 
nomena can be formulated. Many natural laws, especially those con- 
cerned with rates of change, can be phrased as equations involving deriva- 
tives or differentials. For example, when a verbal statement of Newton’s 
second law of motion is translated into mathematical symbols, there 
results an equation relating time derivatives of displacements to forces. 
A study of such equations then provides a complete qualitative and 
quantitative characterization of the behavior of mechanical systems under 
the action of forces. Several broad types of equations studied in this 
book characterize physical situations of great diversity and practical 
interest. 

The first half of this chapter is concerned with preliminaries and special 
techniques devised for the solution of the first-order equations arising 
commonly in applications. The second half contains a comprehensive 
treatment of linear differential equations with constant coefficients and 
an introduction to linear equations with variable coefficients. Linear 
equations occupy a prominent place in the study of the response of elastic 
structures to impressed forces and in the analysis of electrical circuits and 
servomechanisms. They also appear in numerous boundary-value problems 
in the theory of diffusion and heat flow, in quantum mechanics and fluid 
mechanics, and in electromagnetic theory. 


PRELIMINARY REMARKS AND ORIENTATION 

1. Definition of Terms and Generalities. Any function containing var- 
iables and their derivatives (or differentials) is called a differential expres- 
sion, and every equation involving differential expressions is called a 
differential equation. Differential equations are divided into two classes, 
ordinary and partial The former contain only one independent variable 



6 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 


and derivatives with respect to it. The latter contain more than one 
independent variable. 

The order of the highest derivative contained in a differential equation 
is called the order of the differential equation. Thus 



+ 5y 2 * 0 


is an ordinary differential equation of order 2, and 



d 2 y 
dx dt 


+ yxt 


0 


is a partial differential equation of order 3. 

A function y = <p(x) is said to be a solution of the differential equation 

F(x,V,V') = 0 , ( 1 - 1 ) 

if, on the substitution of y ~ <p(x) and y f — v(x) in the left-hand member 
of (1-1), the latter vanishes identically. 1 Again, y = <p(x) is a solution 
of the second-order equation F(xjjy/,y") = 0 when the substitution 
y = <p(x) f y f = <p\x), y" — <p”(x) reduces this to an identity in x. Simi- 
larly for equations of order n . 

For example, the first-order differential equation 

?/ + 2 xy - <-~* S = 0 (1-2) 

2 f 2 
has a solution y = xe T , because the substitution of y — xe T and //' - 

e~ x — 2x 2 e~~ iX in (1-2) reduces it to an identity 0^0. Also, the equation 

y" + y = 0 

has a solution y — sin x, as can be easily verified by substitution. 

We begin our study of differential equations with the first-order equation 
(1-1), which w r e suppose can be solved for y f to yield the equation 

y' = ( 1 - 3 ) 

For reasons which will become clear presently, we shall always assume 
that f(x,y) is a continuous function throughout some region in the xy 
plane, and we shall study the solutions of (1-3) [or, equivalently, of (1-1)) 
in that region. 

The geometrical meaning of the term solution of (1-3) is suggested at 
once by the interpretation of the derivative y f as the slope of the tangent 
line to some curve y = tp(x ), for if (x,y) is a point on the curve y = <p(x), 

1 Here, as elsewhere in this book, primes are used to denote differentiation: y ' ss dy/dx, 
y” m d^y/dx 2 , . . j/ (w) * d n y/dx n . 



SEC. 1 ] PRELIMINARY REMARKS AND ORIENTATION 7 

and if at every point of this curve the slope is equal to f(x,y) f then (p(x) is 
a solution of (3-3). 

One can get an idea of the shape of the curve y * <p(x) in the following 
way: I^et us choose a point (x 0 ,yo) and compute 

y' - f(x 0 ,y 0 ). ( 1 - 4 ) 

The number /(x 0> yo) determines a direction of the curve at (x 0 ,?/o)- Now, 
let (x lf 7/j) be a point near (x 0 ,2/o) in the direction specified by (1-4). Then 
\j — f{x\,y\) determines a new direction at (xi,yi) (Fig. 1). Upon proceed- 
ing a short distance in this new 
direction, we select a new point 
(x 2 , 1 / 2 ) and at this point determine 
a new slope ?/' — f(x 2 ,y 2 )- As this 
process is continued, a curve is built 
up consisting of short line segments. 

If the successive points (x 0 ,//o), 

(*i,y 1 ). (xn.Uz), ■■■, (zn,Vn) are 
chosen near one another, the series 
of straight-line segments approxi- 
mates a smooth curve y = <p(x) 
which is a solution of (1-3) associ- 
ated with the choice of the initial point (xod/o)- A different choice of the 
initial point will, in general, give a different curve, so that the solutions of 
Eq. (1-3) can be viewed as being given by a whole family of curves. Such 
curves are called integral curves , and each curve in the family represents 
a particular solution or an integral of our equation. 

Also, we can make a surmise that, unless f(x,y) in the right-hand member 
of (1-3) is a badly behaving function, for each choice of the initial point 
there will be just one solution of Eq. (1-3). This surmise is capable of 
proof, which we do not give here because it requires the use of analytical 
tools which are not provided in the usual calculus courses. However, the 
statement of essential facts is easy to grasp, and since it will facilitate the 
understanding of subsequent developments, we give it here as a basic 
theorem. 

Existence and Uniqueness Theorem. The equation y* — f{x,y) has 
one and only one integral curve passing through each point of the region in 
which hothf(x r y) and df/dy are continuous functions. 1 

Unless a statement to the contrary is made, we shall suppose that the 
restrictions imposed on f(x f y) in this theorem are fulfilled, so that Eq. 

1 It suffices to suppose that \df/dy\ is bounded in the region. Proofs of this theorem 
are contained in many books on differential equations, for example, E. L. Ince, “Ordi- 
nary Differential Equations/' p. 62. See also See. 17 of this chapter. 




8 ORDINARY DIFFERENTIAL EQUATIONS {CHAP. 1 

(1-8) has a unique solution for each choice of (xo,yo) in the appropriate 
region of the xy plane. 

Since by changing the initial value y\ x » y(x 0 ) we get a family of 
curves depending on the arbitrarily chosen value y(x o), the equation of 
this family can be written in the form 

y = v(x,c ) (1-5) 

involving one arbitrary constant c, corresponding to the arbitrary choices 
of y(x o). A particular curve of the family (1-5) passing through (x 0 ,y 0 ) 
is then determined by the value of c such that y Q = <p(x 0) c). 

A solution of the first-order equation (1-3) involving one arbitrary 
constant is called a general solution } Such solutions are often written in 
the implicit form 

$(x,y,c) = 0, (1-6) 

where it is understood that (1-6) can be solved for y to yield the explicit 
form (1-5). In practice it may not be necessary to exhibit the explicit 
form. The essential feature of the general solution [be it given by (1-5) 
or (1-6)] is that the constant c in it can be determined so that an integral 
curve passes through a given point (x 0 ,y 0 ) of the region under consideral ion. 

We illustrate this 1 ) 3 ' demonstrating that throughout the xy plane the 
general solution of Eq, (1-2) can be written as 

y = e~~ x \x + c). (1-7) 

The fact that (1-7) is, indeed, a solution is easily verified by substituting 
(1-7) in (1-2). Moreover, it is a general solution, because on setting 
x = x 0 and y = yo we get 

2/o = e~~*%(xo + c). (1-8) 

Thus the integral curve passing through (:r 0 ,?/o) corresponds to 

c = y 0 e x o - Xo. 

As another example consider the equation 

dy 

ax 

where f(x) is any continuous function. A general solution of this equation, 
obtained by direct integration, is 

y « ff(x) dx + c. (1-10) 

1 Some first-order equations may have solutions which cannot be determined from 
the general solution for any value of c. Such solutions, called singular solutions^ arise 
only when the conditions imposed on f(x,y) in the basic theorem are not fulfilled. 



SEC. 1] PRELIMINARY REMARKS AND ORIENTATION 0 

We show next that (1-10) is a general solution of (1-9). We denote an 

indefinite integral in (1-10) by F(x), so that dF/dx » f(x). Then (1-10) 
is the same as 

y m F(x) + c. (1-11) 

On setting x = x Q> y = y 0> we get 

Vo « F{x 0 ) + c, 
so that c = y Q — F{x$), 

and we can, therefore, write (1-11) as 

y = Fix) - F(z 0 ) + 2/o 

* F(x)|^ + 2/ 0 . (1-12) 

But from the fundamental theorem of integral calculus, 

f X f(x)dx~F(i) |* 

J Xq 

and therefore (1-12) yields the desired particular solution 

V = / /(a - ) + J/o, (1-13) 

•'XQ 

corresponding to the choice of the initial point (x Q ,y 0 ). 

Formula (1-13) illustrates the procedure of deducing particular solutions 
by integrating the given equation (1-9) between limits. It is frequently 
simpler than the procedure of determining the desired solution by calculat- 
ing the constant c in the general solution from the initial data. 

The foregoing discussion can be extended to equations of higher order. 
Thus, the nth-order equation 

V{x,y,y' y ln) ) = 0, (1-14) 

which we shall write in the form solved for i/ (n) as 

y M „ (1-15) 

has a unique solution for n arbitrarily assigned initial values, 

y{x 0 ), y'{x 0 ), . . . , y {n ~ l) (x 0 ), (1-16) 

whenever the function f in (1-15) is continuous together with the partied 
derivatives df/dy y df/dy f , . . df/dy (n ~^ l) . 

When the values in (1-16) are varied, we get a family of curves, the so- 
called n-parameter family , corresponding to n independent choices of 
constants in (1-16). The equation of this family of solutions can be written 
in the form 


y « ?(x,Ci,C 2 ,. . 


(1-17) 



10 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 

involving n arbitrary constants c % . A solution such as (1-17) is called a 
general solution of the nth-order equation (1-15) [or (1-14)], provided that 
the constants c* in (1-17) can be determined for every given set of arbitrarily 
assigned initial values (1-16). The general solution (1-17) may also appear 
in an implicit form as 

<f>(x,!/,Ci,C2,---,Cn) = 0, (1-18) 

which on solving for y should give (1-17). 

The meaning of the initial conditions (1-16), as they bear on the unique- 
ness of solution of the second-order equation 1 F(x,y,y\y n ) = 0, is that the 
integral curve of this equation is determined at x « x 0 if the ordinate 
2/o = y(zo) and the slope y'{x 0 ) are specified. 

To determine uniquely the solution of the third-order equation, we must 
specify the value of the ordinate 2 /q, the slope y Q , and the value of the 
second derivative yl at x = x 0 . 

In the following nine sections we shall deal with first-order equations, 
which we can write in the differential notation as 

P(x,y) dx + Q(x,y) dy = 0. (1-19) 

If Q(x,y) it 0, Eq. (1-19) gives 

dy _ P(x,V ) 
dx Q(x,y) 

which is in the form (1-3) with f(x,y) = —P(x,y)/Q(x,y). 


PROBLEMS 


Classify the following differential equations as ordinary or partial, and determine 
their orders: 


1 . 


dx 4 \dx) 


3. y' -f sin y + x ® 0; 

5. y " -f- x 2 y' + xy *» sin x; 


7 . Vy + Vjr - y’; 


2 . 


d 4 z 
dx 4 


+ 2 

dx dy 


d 4 z 
dy 4 * 


4. dy « V^l - j/ 2 dx ; 



8. y' A + y' » I/"'. 


Verify that the given expression is a solution of the given differential equation: 

9. y ce x , y' « y; 

10. 2e v - e* 4- ce“* y' - e*~ v - 1; 

11. y Ci sin x + 02 cos x, y" 4- y ■» 0; 

12. y *» Ci sinh x 4" cs cosh x, y" — y * 0; 

18. xy » Jf(x) dx y xy' 4* y » /(x). 

14. Integrate y' » 2x to show that its general solution is a family of parabolas y m x* 
4- c. Determine integral curves of this equation through (0,0), (1,1), (0,1), (1,-1). 



SBC. 2J PRELIMINARY REMARKS AND ORIENTATION 11 

IB. Determine the integral curve for y" » 2i such that j/(0) - 0 and y'(0) — 1. What 
is the general solution of this equation? 

2. The Slipping of a Belt on a Pulley. To illustrate the prominence of 
differential equations in the study of various phenomena, this and the 
following three sections are primarily concerned with the task of setting 
up differential equations from physical principles. 1 Such solutions as 
are included are intended merely as a preview of the systematic discussion 
given in the subsequent sections. If he wishes, the reader may confine 
his attention to the derivation of the equations only and return to the 
question of solution after this systematic discussion has been assimilated. 

The first, example is given by the 
bell -pulley arrangement of Fig. 2, 
which is now to be analyzed Con- 
sider an element of the belt, of length 
As, which has end points Fund Q and 
subtends an angle AO at the center 
0. Let T be the tension at P and 
T + AT at Q, and let A F be the 
normal component of force on As 
due to the pulley. Thus A F is the 
component, along the radius ON, 
of the total resultant, force exclu- 
sive of T and T + AT. 

Assume that the belt is stationary 
and that the pulley rotates, so that 
there is slipping. Since the element 
As is in static equilibrium, the components of force along ON must 
balance. This gives 

A 0 AO 

(T + AT) sin b T sin — - AF, (2-1) 

2 2 

provided the weight of the belt is negligible or provided the pulley axis is 
vertical. Equating forces at right angles to ON leads to 

AO AO 

(T + AT) cos T cos — = g AF, (2-2) 

2 2 

where ju is the coefficient of sliding friction. 2 From (2-2) we may deduce 
A T ~ » AF, AO 0, (2-3) 

1 Further problems of the sort are treated in Sees. 18 to 20. 

* We define n by (2-2) and regard it as an experimental fact that ^ approaches the co- 
efficient of friction for flat surfaces or, at any rate, some limit independent of 6 as 
AO 0. 


N 




12 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 

where the symbol ~ (read “is asymptotic to”) means 1 that the ratio of 
the quantities on each side tends to 1. Thus, if a ~ b, then lim a/b « 1, 
Equations (2-1) and (2-3) together show that AT — ► 0 as Ad — ► 0. Since 
sin (A6/2) ~ AB/2 } Eq. (2-1) now gives 

T AS ~ A F. (2-4) 

Dividing (2-3) by (2-4) leads to AT/ (T Ad) ~ /x, which becomes 


dT 

¥d$ 




(2-5) 


since lim (A T/AB) = dT/dS . 

Separating the variables in (2-5) yields dT/T = n dB y which, upon in- 
tegration, becomes log T « pd + c. The initial condition T == T 0 when 
B « 0 gives c = log T 0 , so that, taking exponentials, 

T « 7V M *. (2-6) 


PROBLEMS 

1. Obtain Eq. (2-3) by equating torques about the point 0. 

2. If the pulley axis is horizontal, and if the belt weighs w lb per ft, show that Eq. 
(2-3) becomes n AF ~ AT — w As cos 0 and Eq. (2-4) becomes A F ^ T Ad -f- u> As sin 0, 
where A F is the normal component of the reaction of the pulley on As and the line OA 
in Fig. 2 is horizontal, with P above it. Deduce the differential equation dT/dd — nT - 
wr{p sin 9 + cos $), where r is the radius. 

3. Show that the equation in Prob. 2 becomes d(Te~^) m wre"** 9 (ju sin 6 -f- cos 9) dd 
when multiplied by and thus obtain the solution. 

8. Growth. Equation (2 -5), which was obtained for the tension in a 
dipping belt, arises in many other connections. For example, radium de- 
composes at a rate proportional to the amount present. If this amount 
is A at time t , the foregoing statement means 

dA 

-=-kA, k> 0, (3-1) 

at 

the negative sign being chosen because A decreases as t increases. A 
similar equation is followed by the growth of populations in certain cir- 
cumstances. Thus, the rate of increase being nearly proportional to the 
number N present, one can write dN/dt = kN. Again, certain organisms 

1 The relation symbolized by ~ has many of the properties of strict equality. For 
example, if a ~ b and b ~ c then a ~ c. To see this, observe that a/b — > 1, since 
a ~ b; and b/c X, since b ~ c, and hence, by multiplication, (a/b)(b/c) 1 *1. Thus 

a/e — ► 1, which is to say, a ~ c. The reader may verify similarly that a ~b and 
c ~d together imply ac ~bd and a/c ~ b/d. Finally, if a ~ b and b is constant, we 
may write lim a * 6. These properties are freely used in the text. 



PRELIMINARY REMARKS AND ORIENTATION 


13 


SEC. 3 ] 

grow at a rate proportional to their size S at a given time so that dS/dt =» 
fcS. 


Example 1. In a colony of bacteria each bacterium divides into two after a time inter- 
val, on the average, of length r . If there are n bacteria at time t - 0 and m at time t « 1, 
with n large, find the approximate value of r. 

The hypothesis implies that dN/dt « kN, approximately, with greater and greater 
accuracy as the number of bacteria N becomes large. Separating variables gives dN /N 
*» kdt. Now t *» 0 corresponds to N n, and t «* 1 corresponds to N ** m, by hy- 
pothesis. Thus, 



(3-2) 


and similarly 



(3-3) 


since N doubles in the interval r. Equation (3-2) gives log m — log n k, and (3-3) 
gives log 2 «* kr f so that 

log 2 

T ** 

log m — log n 


This problem illustrates the useful method of integration between limits for the determina- 
tion of constants. A justification of this procedure is implicit in Sec. 1, Eq. (1-13). 

Example 2. A radioactive substance A decomposes into a new substance B, which 
in turn decomposes into a third substance C. Set up a differential equation for the 
amount of B at time t. 

The rate of increase of B is equal to the rate at which B is formed from A minus the 
rate at which B decomposes. Thus, denoting the amounts by A and £, 


dB 

dt 


dA 

dt 


- hB . 


(3-4) 


This equation has two unknowns, A and B. By (3-1), however, A ** ce~* l t so that 
(3*4) becomes 

dB 

* kce~ kt — kiB. (3-5) 

dt 


A method of solving (3-5) is given in Sec. 10. 


PROBLEMS 

1. If 3 g of a radioactive substance is present at time t *» 1 and 1 g at t =* 4, how much 
was present initially? 

2. In Example 2 of the text set up the differential equation for the amount of substance 
C present at time t. 

3* By actual substitution, determine a and p in such a way that B « ae ** is a solu- 
tion of Eq. (3-5). 

4. The rate of decomposition of a certain chemical substance is proportional to the 
amount of the substance still unchanged. If the amount of the substance at the end of 
t hr is x and xo is the initial amount, show that z m where k is the constant of 

proportionality. Find k if x changes from 1,000 to 500 g in 2 hr. 



14 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 

5. A torpedo moving in still water is retarded with a force proportional to the veloc- 
ity. Find the speed at the end of t see and the distance traveled in t sec if the initial 
speed is 30 mph. 

0. The rate at which a body is cooling is proportional to the difference in the tempera- 
tures of the body and the surrounding medium. It is known that the temperature of a 
body fell from 120 to 70°C in 1 hr when it was placed in air at 20°G. How long will it 
take the body to cool to 40°C? 30°C? 20°C? 

7. The percentage of incident light absorbed in passing through a thin layer of material 
is proportional to the thickness of the material. If 1 in. of material reduces the light to 
half its intensity, how much additional material is needed to reduce the intensity to 
one-eighth of its initial value? Obtain the answer by inspection, and check by solving 
an appropriate differential equation. 

4. Diffusion and Chemical Combination. Problems involving chemical 
reactions and the formation of mixtures often lead to differential equations; 
the discussion is similar to that of Sec. 3. For example, suppose that a 
tank contains g gal of water and that brine containing w lb of salt per 
gallon flows into the tank and out again at a constant rate r gpm, starting 
at time t — 0. At the same time a piece of rock salt is dropped into the 
tank, where it dissolves at a constant rate of q lb per min. The mixture 
being kept uniform by stirring, it is required to find the amount of salt 
present at any time t > 0. 

This problem may be taken as the typical problem for many questions 
involving chemical reactions, mixing, and going into solution. The dif- 
ferential equation is obtained by writing down the equation of continuity 
(increase equals income minus outgo) fur the amount of salt. Call this 
amount x » x(t) at time t. In the time interval from t to t + At the 
number of gallons entering the tank is r At, since the rate of flow is r. 
Now each gallon contains w lb of salt. Hence the r At gal contains 

wr At (4-1 ) 

pounds of salt, and this, then, represents income due to the inflowing brine. 
The income due to the dissolving salt is 


q At , (4-2) 

by the definition of q. 

It remains to compute the amount of salt lost in the mixture leaving 
the system. The number of gallons leaving is r At, the concentration of 
the mixture in pounds per gallon is x/g at time t , and hence the number 
of pounds leaving is 


- r At. (4-3) 

Q 


Here £ denotes the mean value of x over the interval ( t , t + At). We 
assume x to be continuous, so that 



SEC. 5] 


PRELIMINARY REMARKS AND ORIENTATION 


15 

(4-4) 


lim x ® x. 
&t -* o 

From (4-1), (4-2), and (4-3) we obtain 


Ax « wr At + q At r At, 

9 


which gives 

dx rx 

— a wr + a 

dt g 

when we divide by At and let At 0, using (4-4). 


(4-5) 


Example: Find the concentration of salt at the end of 4 min when w « 1, g 2 
q *** 3, r ** 4. 

The differential equation is dx/dl * 7 — 2x or &r/(7 — 2x) ** d/. Multiplying by ~2 
and integrating give log (7 — 2x) = —2t -f c. Since x — 0 when t — 0, it is necessary 
that c = log 7, so that ~~2t «* log (7 — 2x) — log 7 * log (1 ~ 2x/7) or, taking exponen- 
tials, 1 - 2x/7 ** c" 2 *. This gives the amount of dissolved salt x at the end of t min. 
Putting t «= 4, solving for x , and noting that the concentration is not x but x/g give 
J4O - e" 8 ) as the final answer. 


PROBLEMS 

1. Solve the example of the text by the method of integration between limits. (See 
Example 1, See. 3. Here x = 0at< = 0, x=*xat^=»4) 

2. How would the discussion in Sec 4 change if the rock salt had been added at time 
t «■ io instead of time l ■» 0? 

3 . How would the discussion in Sec 4 change if the rock salt dissolved at a rate pro- 
portional to the amount undissolved, rather than at the constant rate q‘l Hint ■ If A is 
this amount, dA/dt ** — kA . From this find A at time t , and from that find q * — dA /dt. 

4 . Let A be the amount of a substance at the beginning of a chemical reaction, and let 
x be the amount of the substance entered in the reaction after t sec. Assuming that the 
rate of change of the substance is proportional to the amount remaining, deduce that 
dx/dt » c( A — x), where c is a constant depending on the reaction. Show that x ® 
A{ 1 - e" cl ). 

6. Ixd. a solution contain tw r o substances w r hose amounts expressed in gram molecules, 
at the beginning of a reaction, are A and B. If an equal amount x of both substances 
has changed at the time t f and if the rate of change is jointly proportional to the amounts 
of the substances remaining, obtain the equation dx/dt = fc(A — x){B — x). Solve, 
assuming that x ** 0 when t *= 0. 

6. Formulate the appropriate differential equation if the rate at which a, substance 
dissolves is jointly proportional to the amount present and to the difference between the 
actual concentration and the saturate concentration. 

5. The Elastic Curve. Consider a horizontal elastic beam under the 
action of vertical loads. It is assumed that all the forces acting on the 
beam lie in a plane containing the central axis of the beam. Choose the 
x axis along the central axis of the beam in undeformed state and the posi- 



16 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 


tive y axis down (Fig. 3). Under the action of external forces Fi the beam 
is bent and its central axis deformed. The deformed central axis, shown 
in the figure by the dashed line, is known as the elastic curve, and it is an 
important problem in the theory of elasticity to determine its shape. 

A beam made of elastic material that obeys 
F, *2 F 3 Hooke’s law is known to deform in such a way 

that the curvature K of the elastic curve is 
proportional to the bending moment M. In 
fact, 

y" M 

K — — (5-1) 

[1 + (y') 2 ]* El 

Fig. 3 where E is Young’s modulus, I is the moment of 



inertia of the cross section of the beam about a 
horizontal line passing through the centroid of the section and lying in 
the plane of the cross section, and y is the ordinate of the elastic curve. 
The important relation (5-1) bears the name Bernoulli-Euler law . When the 
deflection of the beam is small, the slope of the elastic curve is also generally 
small and one can neglect the term ( y ') 2 in (5-1) to obtain an approxi- 
mate equation 


y 


tt 


M 

~ei 


(5-2) 


The bending moment M in any cross section of the beam is equal to 
the algebraic sum of the moments of all the forces F t acting on one side 
of the section. The moments of the forces F t are taken about a horizontal 
line lying in the cross section in question. 

Example: Consider a cantilever beam of length l t built in at the end x «* 0 and carry- 
ing in addition to a distributed load ic(x) lb per ft a concentrated load W lb and a couple 
L ft>-lb applied at the end x *= l (Fig. 4). 

The resultant moment in a cross section x ft from the end x *■ 0, produced by the 
loads acting to the nght of that section, is 

Mil) - - X)w(i) d£ + W(l - x) + L. (5-3) 



Fio. 4 



THE SOLUTION OS' FIRST-ORDER EQUATIONS 


17 


SEC. 6] 

If w(x) m 0 and L «* 0, this formula yields M m W(l — x)> and hence, from (5-2), the 
differential equation of the central line of a cantilever beam subjected to the end load If is 




On integrating this equation we get 


r \2 6 ) 


+ CiX + Oj. 


The integration constants Oj and c -2 can be evaluated from the conditions y( 0) * 0, 
y'(Q) « 0, stating that the displacement and the slope of the central line vanish at the 
built-in end. It is readily checked that these conditions lead to 


(*•-?> 


so that the displacement d at the free end is d * WP/ZEI. 


PROBLEMS 

1. A beam of length l is freely supported at its ends and is loaded in the center by a 
concentrated ve*rtical load IT, which is large in comparison with the weight of the beam 
(see Fig. 5). By symmetry, the behavior of tins beam is the same as that of a cantilever 
beam of length 1/2 loaded by a concentrated 

load of magnitude If/2 at its free end. Verify W 

this equivalence by direct computation of the 
elastic curve. Hint: 

M ** — 0<x<^> | SSVs> * ■- — -- — ”” 

w w 

W l 2 2 

- - -- (J - x ) » -<x <1 FlG. 5 

2. A uniform unloaded beam of length l weighs w lb per ft. Find the maximum de- 
flection when it is used as a cantilever beam and also when it is freely supported at each 
end. Hint: Since the reaction at the end x ® l is R » If/2, the moment in the cross 
section at a distance x from the end x » 0 is 


AT - wj (£ - x) d£ - ~ (1 


THE SOLUTION OF FIRST-ORDER EQUATIONS 

6. Equations with Separable Variables. Generally speaking, the prob- 
lem of solving differential equations is a very difficult one. Even such a 
simple equation as y' = f(x f y) cannot be solved in general; that is, no 
formulas are available for solving the general differential equation of the 
first order. It is possible, however, to classify some of the first-order dif- 



ORDINARY DIFFERENTIAL EQUATIONS 


18 


[CHAP. 1 


ferential equations according to several types and to indicate special 
methods of solution suitable for each of these. 

Prominent among these types are the equations with separable variables f 
that is, equations which can be put in the form 


P(x) dx + Q{y) dy * 0, 


where P(x) is a function of x only and Q(y) is a function of y only. This 
type of equation has already been encountered in the special examples 
solved above. Its general solution is 


Jp(x) dx + J Q(y ) dy = c, 


where c is an arbitrary constant. In order to obtain an explicit solution 
all that is necessary is to perform the indicated integrations. 

Example: Find a solution of y 4- c r y * e r y 2 which goes through (0, V*j). 

The equation can be written as y f -f- e z (y — y 2 ) * 0 or 


Integration gives 



e x dx *= 0. 


log h e r « c, 

1 - y 


which is a general solution. Putting x » 0, y = gives c == log 1 -f- e° » 1, so that the 
required particular solution is 

log h * 1. 

I - y 


PROBLEMS 

Solve the following differential equations. In Probs. 4 to 6 find a solution through the 
point (0,1). 

1. Vl — x 1 dy * VY — y 2 dx. 2. y ' = xy 2 — x. 

« , sin2 x A ■ 2 j 2 ^ 

3. y « — 4. sin x cos^ y dx « cos^ x dy. 

sin y 

5. VTf x dy « (1 -f I/ 2 ) dx. 6 . y' » • 

1 +z 

7. Homogeneous Differential Equations. A function /(x,y) of the two 

variables x and y is said to be homogeneous of degree n provided that 

f(Xx,Xy) a \ n f(x,y), X > 0. 

Thus, f(z,y) = x 3 4* + y 3 is a homogeneous function of degree 3, 

and /Or, 2 /) = x 2 sin (x/y) + xy is a homogeneous function of degree 2, as 
follows at once on replacing x by Xx and y by Xy. 

If the differential equation is of the form 



SBC, 7] THE SOLUTION OF FIRST-ORDER EQUATIONS 19 

P{x,y) dx + Q(x f y) dy * 0, (7-1) 

where P(x } y) and Q(z,y) are homogeneous functions of the same degree, 
then (7-1) can be written in the form 


y' ~ 


P(*,y) 

Q(x,y) 


s <t>(*,y), 


(7-2) 


where 0(x,y) is a homogeneous function of degree zero; that is, 
<t>(\x,\y) s X°^(x,2/) as 0(x,y). 

If X is set equal to l/x, then 


<t>(x,y) s <#»(Xx,Xy) 



which shows that a homogeneous function of degree zero can always be 
expressed as a function of y/x . This suggests making the substitution 
y/x = v. Then, since y = rx, 


dy 

dx 


dr 

— x + v. 
dx 


Substituting this value of dy/dx in (7-2) gives 

dv 

X — -f V = 

dx 


This equation is of the type considered in Sec. 6. Separating the variables 
leads to 

dv dx 

0(1, v) — v x 

which can be integrated at once. 


Example: Solve 


y* -f- x 


dy 

dx 


xy 


dy 

dx 


This equation can be put in the form 
dy y^_ 

dx 


y 2 m {y/x) 2 

xy - x 2 y/x - 1 
Letting y/x « v and computing dy/dx from y ® vx give 


p + i 


dv 

dx 


v * 

v ^T 


dv 


dx v — 1 


Separation of the variables leads to 

dx 1 — v 


dv < 



ORBINARY DIFFERENTIAL EQUATIONS 


20 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

and integration yields log x -f log v — v — c or log vx — v m c. Since v m y/x, the final 
answer is log y — y/z ■» c. 

PROBLEMS 

Solve the following differential equations. In Probs. 4 to 6 find a solution through 
the point (1,1). 


1 . ( x 2 -f y 2 ) dy 4- 2 xy dx * 0. 

ydy y 

3. x cos -—-=*=?/ cos x. 

XOX X 

5 * x 2 y dx » (x 3 — y 3 ) dy. 


2. xy' - y Vx* — y 2 

4. (x -f y)y' ** x -y. 

- dy xy - y 2 


Some of the following equations are separable; some are homogeneous. Solve them. 


7. sinh x dy + cosh y dx =» 0. 

9. x(V7y + y)dx » x 2 dy. 
11. xy' ** y + xe v/a! . 


dx x - Vxy 
10. x 2 y' — y 2 ■» x 2 yy'. 

12. y' =* y' log y + tan x see 2 x. 


8. Exact Differential Equations. An expression P(x,y) dx + Q(x,y) dy 
is said to be exact if it coincides with the differential 

dF dF 

dF = — dx H dy 

dx dy 

of some function F(x,y), that is, if 

dF dF . 

P(x,y) dx + Q(x,y) dy = — dx dy. (8-1) 

dx dy 

In these circumstances the equation 

P(x,y) dx + Q(x,y) dy = 0 (8-2) 

is simply dF = 0, and its general solution, therefore, is 

F{x,y) « c. (8-3) 

When a function F(x,y) satisfying the relation (8-1) exists, we conclude 
that 

dF dF 

— * P(x,y), — = Q(x,y). (8-4) 

dx dy 

Moreover, if d 2 F/(dx dy) = d 2 F/(dy dx), we obtain by differentiating (8-4) 
a necessary condition, 

dP dQ 

— « — ( 8 - 5 ) 

dy dx 

for the existence of F(x,y). This condition also suffices to construct F(x,y) 
in every rectangular region throughout which P, dP/dy, and dQ/dx are 



SEC* 8] THE SOLUTION OP FIRST-ORDER EQUATIONS 21 

continuous. 1 Indeed, on integrating the first of Eqs. (8-4) with respect to 
x , we get 

=* / P(x,v) dx + f(y), ( 8 - 6 ) 


where f(y) is an arbitrary differentiable function of y, since y appearing 
in the integral of (8-6) is treated as a constant. We next determine f(y) 
so as to satisfy the second of Eqs. (8-4). Differentiating (8-6) with respect 
to y and equating the result to Q(x>y) give 


so that 

This determines 


dF d 

— - - PM dx+f'(y) = Q(x,y), 
dy dy J 

f'iv) - Q(x,y) - ■— / P{x,y) dx. 
dy 1 

f(y) = / 1 Q(x,y) ~ ^ / P(x,y) dx] dy, 


(8-7) 


provided the expression in the brackets in (8-7) is a function of y only. 
But that is always the case, since its derivative with respect to x is dQ/dx — 
dP/dy, and this vanishes whenever (8-5) holds Accordingly, the substitu- 
tion of f(y) from (8-7) in (8-6) gives the function F(x,y) and thus the de- 
sired solution F{x,y) — c. 


Example: Solve the equation 

(2 xy + 1 )dx + ( x 2 -f 4 y) dy - 0. 


This equation is exact, since dP/dy *= dQ/dx ~ 2x. Thus there is a function F(x t y) such 
that 


dF 

dx 


2 xy 4- 1, 


OF 

dy 


x 2 -f 4y. 


( 8 - 8 ) 


From the first of Eqs. (8-8) we conclude that 

F(x,y) - j(2xy + 1) dx +f(y) 


- x 2 y 4- x 4-/(y). (8-9) 

To satisfy the second of Eqs. (8-8), we must have 
d P 1 

— •• ** +/'(y) - i ! + 4» 
so that f\y ) ** 4y. 


The integration yields 

/(If) - V, 

and the substitution in (8-9) gives F(x,j/) « z 2 i/ 4* ^ + 2y J . The desired solution, there- 
fore, is 

x f y 4* x + 2y* « c. 


1 For details and general discussion see Chap. 5, Sec. 9. 



22 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP, 1 


PROBLEMS 


Integrate the following equations if they are exact: 

1. («* *f 3) dx + dy «* 0; 2. ( 2x -f - e xl *) dx « \ e xtv dy; 

\ V / 2 

5. (3x 2 y ~ y z ) (h ** (3y 2 x - x 3 ) dy; 4. x dy + y dx » 0; 

V V 1 y 

6. -*5 cos - dx * - cos - dy; 6. x dx *f y dy «* 0; 

x* x x x 

7. (3 x 2 y - y 3 ) dx - (x 3 + 3y*i) dy - 0; 

8. (y cos xy + 2x) dx -f x cos xy dy ~ 0; 

9. (y 2 + 2xy + 1) dx + (2xy + x 2 ) dy - 0; 

10. 3 x 2 y dx + (x 3 -- 3y 2 x 2 ) dy - 0. 

9. Integrating Factors. Suppose that 


has a solution 


M(x,y) dx + N(x,y) dy — 0 
F{x,y) = c, 


(9-1) 

(9-2) 


where F(x,y) is a differentiable function. On differentiating (9-2) with 
respect to x , we get 

dF dF 


dx 


+ —</' = o, 

dy 


(9-3) 


and from (9-1) we find 


M(x,y) + N{x } y)y' = 0. 


(9-4) 


The elimination of y ' from (9-3) and (9-4) gives 
dF/dx dF/dy 


M{x,y) N (x,y) 


M(x,y), 


(9-5) 


where is the value of the common ratio. It follows from (9-5) that 

dF dF 

— = M(ar,y)JI/ («,y), — = n(x,y)N(x,y) 

dx dy 

and hence that 

y(x,y)(M dx + N dy) = 0 


is an exact equation; namely, it is the equation dF ~ 0. 

The function y(x f y) is termed an integrating factor. It is clear from the 
above discussion that every equation (9-L) has an integrating factor and, 
in fact, an unlimited number of them. 1 Nevertheless, it must not be con* 
eluded that an integrating factor can always be found easily. In simpler 
cases, however, it can be found by inspection. 

1 Some integrating factors introduce extraneous solutions y which make y(x,y) « 0 but 
do not satisfy (9-1). 



SEC. 10] THE SOLUTION OF FIRST-ORDER EQUATIONS 23 

Thus, in order to solve 

xdy — ydx = 0 

which is not exact as it stands, multiply both sides by l/xy . Then the 
equation becomes 

dy dx 

= 0 , 

y x 


which is exact. Another integrating factor for this same equation is l/x % . 
Similarly, multiplication by 1 /y 2 makes the equation exact. 

Example: Solve the differentia! equation 

(y 2 — x 2 ) dy 4- 2 xy dx « 0. 

This is not an exact equation, hut on rearrangement it becomes 
y 2 dy 4- 2xy dx — x 2 dy « 0, 


which can be made exact with the aid of the integrating factor 1/y*. 
equation is 

2ri/ dx - x 2 dy 
dy + - t 0, 

which integrates to 

1/4 — = c. 

y 


The resulting 


PROBLEMS 

The following problems give a few of the integrable combinations that commonly oc- 
cur in practice. Verify the equations by differentiating: 


4 , (. _i A xdy - ydx 

1. d (Un J - + yi , 

8 . d (f ) = 

\y / y“ 

6. d(x 2 4- y 2 ) * xdx 4- y dy; 


' t . d ( log y)„ x Jv^i d * 

\ x/ xy 

4 . d (v\ = z*l^s± 

\x/ X 2 

6. d{xy) * x dy 4- y dx. 


x dy — ydx 


Solve the following equations by finding a suitable integrating factor: 


7 . x dy 4- x 2 dx — y dx; 
9 . xdy -j- Zy dx * xy dy; 
11. xdy — ydx « xy dy; 


8. (xy 2 4 -y)dx =* (x 2 y - x) dy; 
10. (x 2 4- y 2 4- 2x) dy « 2y dx; 
12. (x 2 - y 2 ) dy * 2xy dx. 


10. The First-order Linear Equation. An equation of the form 


+ M(x)y = N(x) 


is termed linear for reasons given in Sec. 21. 



24 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

If we set y » uv, where u and v are functions of $ to be determined later, 
we get on substitution in (KM) 

uv f + vu f + Muv * N, 

or v(u' + Mu) + uxf — N. (10-2) 

If u is suitably chosen, the parenthesis in (10-2) can be made equal to 
zero, thus reducing (10-2) to a simpler form. To this end, set 

u ' + Mu = 0, (10-3) 

which is a separable equation for u. We get 

du 

b M dx = 0, 

so that log u + J M dx = c. (10-4) 

Since any solution of (10-3) reduces (10-2) to the form 

uv' » A r , (10-5) 

we choose the simplest one, corresponding to c = 0. With this choice, 
(10-4) yields 

u = e-f v ‘ u , ( 10 - 0 ) 

and (10-5) becomes 

v' = NeJ M d \ (10-7) 


Since the right-hand member in (10-7) depend", only on x, we get, on in- 
tegrating, 


: fNef" 1 


’ dx + r. 


Kecalling the assumption that // = uv, we get the general solution 

y = <>-/" d 'j Nef M dr di + ce~S M dx . ( 10 - 8 ) 

Example I. Solve y' + yvoax * un 2x. Here M(x) ** cosx and N(x) sm 2x. 
Since Jm dx =* Jqoaxdx ** sih x, (10-$) yields 

y zm € ~ Bl » r J * sin 2x dx re““ Bm *, 

which is easily evaluated by replacing sm 2x by 2 sin x cos x. 

Example 2. Solve (x -f 1 )//' *f 2 y » (j + l) 4 . Dividing by x + 1 shows that this 
equation is linear with M 2 f(x 4- 1) and N « (x + l) 3 . Henre 

i*2 dx 

f f xt dx - c J *+i - c 21 ® (x+1) =(i + l) 2 , 
r~/' v = ( r f I)" 2 . 


ulule 



25 


BBC* 11J THE SOLUTION OF FIRST-ORDER EQUATIONS 

Thus (10-8) yields 

y - (x + l)-*f (x + l fdx + c(x + i)~ 3 


(x + l) 4 
6 


+ c(x *f 1)~ 2 . 


PROBLEMS 


Solve the following equations. In Probs. 3 to 5 find a solution through the point 

( 0 ,- 1 ). 

1. (1 + x 2 ) dy “ - x l /) dx. 2. (z 2 + 1)2/' + 2xy » x 2 . 

3. y' * e'** — 2xy. 4 . y' 4- xy — x *= 0. 

6. y r -f y cos x ** cos 8 x. 6, xy' -f V ”* a: 2 sin x. 

7. Show, on writing Eq. (10-1) in the form 

dy + A fy dx — N dx, 

that eJ M dx is an integrating factor of this equation, and thus obtain formula (10-8)* 

Solve the following equations, each of which is separable, homogeneous, exact, or 
linear. (It is instructive to use several methods when possible.) 


8, y' » y 4* cos x — sin x. 
dx 

10. - - + yx * y 
dy 

12. y' 4- yx ■» //. 


q dy ^ y 2 - x/r 2 - y 2 
dx xy 

11. x 2 (l + 4 y z ) dx + 3yx 8 dy - 0. 

13. — — - dx + (1 - e v ) dy » 0. 

V 


11. Equations Solvable for y or y'. Certain special types of equations 
can he solved by writing p ~ dy/dx and expressing p as a function of 
x and y. Another method is to solve for y in terms of x and p and then 
differentiate with respect to x, using dy/dx = p. These procedures change 
the given first-order equation into a new one. 


Example 1. Solve 2 p ? — (2 y 2 4* x)p + xy 2 « 0, where p ■» dy/dx. 

Factoring gives (/> — y 2 )(2p - x) « 0 so that, at each x, we have either p — y* or 
p » x/2. The fact that y is to be differentiable ensures that one or other of these rela- 
tions actually holds throughout an interval. Hence, with p *» dy/dx, they can be re- 
garded as differential equations and solved m the ordinary way. From dy/dx » y 2 
there results 

x 4. 1 * Ch (11-1) 

y 


and from dy/dx * x/2 is obtained 

x 2 

y « - + c 2 . 


( 11 - 2 ) 


These two sets of curves represent the desired solution. Although there is no advantage 
in doing so, one may write (11-1) and (11-2) as a single equation with a single parameter, 



26 


ORDINARY DIFFERENTIAL EQUATIONS 


[chap. 1 


(* + i + c) (v-J + ') -0. 

Example 2. Solve p h vy *» 1, where p — dy/dx. 

Since it is impractical to solve this equation for p to obtain p ** f{y) (which would 
have led to a separable equation), we solve it for y and obtain 


V 



Differentiating (11-3) with respect to x leads to 


which can be written as 
After integration we get 


dy l dp „dp 

dx ^ p* dx ^ dx' 


dx =» — - dp — 4 •©* dp. 



(11-3) 


(11-4) 


which, together with (11-3), gives the desired solution in parametric form. There is no 
advantage in eliminating the parameter p in Eqs. (11-3) and (11-1), even when it is 
possible to do so. Plotting the curves representing the solution as p varies, one obtains 
not only the locus (x,y) but also the slope p at each point. 

The method used to solve the equation in the preceding example can be 
applied to solve the Lagrange equation 

v = xf(y') + g(y'), (n-5) 

where / and g are differentiable functions of y f ~ p. On setting y' = p 
in (11-5) one obtains 

y = *Kv) + y(v)- (n-o) 


Differentiating with respect to x yields 


V 


dp dp 

xf(p)^ + f(p) + g f (p)^ 

dx dx 


which can be written as 


dx ^ f f (p) ^ ^ g'(p) 

dp p ~ f(p) p - f(p) 


(11-7) 


This equation is linear in x; that is, it is of the form dx/dp + M(p)x * N(p), 
and it can be solved by the method of Sec. 10. Its solution for x as a 
function of p, together with (11-6), yields the solution of the original 
equation in parametric form, with p as parameter. 

The reader will find it instructive to apply this method to solve y « 
xy f + (|/') 2 and show that y *** cx + c 2 . 



SEC. 12] 


THE SOLUTION OF FIRST-ORDER EQUATIONS 


27 


PROBLEMS 

Problems 1 and 2 are to be solved by the method of Example 1 ; Probs. 3 and 4 by 
that of Example 2. 

1. p 2 - 2 yp » 3 y 2 . 

3. p 4 « p 2 y -f 2. 

5. p 2 x — p|/ 4" 1 ® 0. 

7. p 2 4* (2x - y)p « 2xp 
3. x « 2p - p*. 

10. Show that Clairaut’s equation y ** xp 4-/(p) is a special case of Lagrange’s equa- 
tion (1 1-5), and thus obtain the solution. 

12. The Method of Substitution. Many first-order equations can be 
solved by a suitable change of variable. This has already been demon- 
strated in the substitution y ~ vx for the homogeneous equation (Sec. 7), 
in the substitution y ~ uv of Sec. 10, and in the use of p = dy/dx as 
independent variable in Sec. 11. Further examples of the substitution 
method are given in this section. 

Thus, the Bernoulli equation 

Vf + P(x)y - Q(x)y * ( 12 - 1 ) 

can be reduced to a linear equation by setting z = 2 / 1 ~ n . 

On dividing (12-1) by y n , we get 

iTV + P(x)y^ x « Q(z). 

But since (y l ~ n Y = (1 — ri)y~ n y', we can write this as 

~ — (y l ~ n Y + P(x)y l ~~ n = Q(x)> 

1 — n 

On making the substitution z — // ~~ n , we get the linear equation 

2 ' + (1 ~ n)/ > (x)2 = (1 — n)Q(x), (12-2) 

which is solvable by the method of Sec. 10. 

The equation 

t ~ '( rxr4r ) <“« 

ax vOjX 4 ~ i>2y ”f* 03 / 

can be solved by the substitution x — w — /i, = p — fc if the constants 

h, k are chosen so as to make the resulting equation homogeneous. This 
procedure, which is simply a translation of axes, is illustrated in Example 2. 

Because of the habitual use of the notation dy/dx , which implies that 
y is a dependent variable, one may fail to recognize that an equation is 
solvable if the roles of x and y are interchanged. For example, an equation 
which is nonlinear in y may become linear if x is regarded as the unknown 
and y is regarded as the independent variable. If an equation seems in- 


2 . p 2 4- 1 - 2 p. 

4 . p 3 4- 2p - e v . 

6. p 2 4- V 2 * 1. 

8 . p 2 4* (x — e*)p ~ xe x . 



28 


ORDINARY DIFFERENTIAL EQUATIONS 


[chap. 1 

tractable as it stands, it is often helpful to interchange x and y , simplify, 
and attempt to solve the new equation. Then interchange x and y in 
the solution of this to obtain the solution to the original equation. The 
procedure, which is illustrated in Example 3, amounts simply to the change 
of variable x « y } y = x. 

Example 1. Solve the equation y’ 4 - y ** xif. 

This is a special case of Bernoulli's equation. Set z =» y~ 2 to obtain z' — 2z *» ~2x 
by direct calculation or by (12-2). The general solution is z * ce 2x 4-x 4 3 2. so that 
j/~ 2 « cc 2 * 4“ is the solution of the original equation. 

Example 2. Solve 

dy x — t/ — 2 
dx x 4* 2/ 4 6 


by means of the substitution x ~ u — h, y ** v ~ k, where h, k are suitably chosen 
constants. 

Substituting gives 

dv ^ u -v ~ (h ~ k 4 2) 
du u 4* v — (h 4- k ~ 6) 


ff h and k are so determined that 


h — k 4 2 * 0, 
h -f k - 6 « 0, 


(12-5) 


then (12-4) becomes the homogeneous equation 


dv u — v 
du u 4- 0 

whose solution is 


w 2 — 2uv — v 2 ** ci 

by Sec. 7. Equations (12-5) give h = 2, A; « 4, so that u » 
A; *■ 4* 4. Substitution in (12-6) leads to the final answer 


x 4 ~ h 


(12-6) 
x 4 2, v * y 4 


x 2 — 2xy — y 2 ~ 4x — 12j/ = c 

after simplification. 

Example 3. Solve (x — f) dy ** y dx. 

Interchanging x and y gives ( y - x 8 ) dx » x dy or y' — y/x * —x 2 . This equation 
is linear in y and gives 2y « cx — x 8 by the method of Sec. 10. Hence the solution of 
the original equation is 2x =* cy — f . 

Example 4. Show how to solve the equation 3/' ** P(ax 4- fy/ 4* c), where a, 6, c 
are constant. 

Let z = ax 4* hy 4* c, so that s' «* a 4 by'. Combining this with the original equa- 
tion gives 2' — a « by' *» 6P(ax 4“ hy 4- c) « 5P(«), or » a 4 6P(z). This equation 
is separable. The procedure fails if b *# 0, but then the original equation is separable. 


PROBLEMS 

Solve the following special cases of Bernoulli’s equation: 

1, f-j~~ 1 — « sm x; 

ax % 


a- »' + v - 



sec. 13] 

3 

8* xy f 4 V 


THE SOLUTION OP FIRST-ORDER EQUATIONS 


29 


i?l + ± mg *. 

%/dx xy® ' 

V 2 log x; 


4. y' - x~*y 4- x~V - 0; 
6. 1/' 4 xy « x 8 y* . 


Reduce the following equations to a form which is homogeneous or has separable 
variables, but do not solve; 


7. y' - 
9. y' - 


x 4 y - l m 
2x 4 y 4" 2 f 
. * — y 4* 2 


8. y'- 
10. y' 


3 x 4 y 4 6 
3x4l/47 ; 

cos (x 4 i/). 


x 4 y 4 3 

Solve by interchanging x and y and using an appropriate method on the result: 
dx 


11 . 08 

dy 

«Vi 


12. y dx - (x 4 y 3 ) dy; 
13, eV; 14. 1 4 xy' tan y * y'. 

Solve the following review problems by any method: 


16. ?/(l 4 x 2 ) 1 dx 4 tan 1 x dy « 0; 


_ dx x . . 

17. 4 - 4 y l 

dy v 


0 ; 


19. e x y' » e* 4 e v ; 

21. dx 4 2x dy « y dy ; 

23. dy - (2y 4 c 3 ') dx; 

26. (x - y 4 1) dx 4 (x 4 V - 1) dy « 

13. Reduction of Order. With y' 

d 

“ - 7, 

dp 


■y 


16. (1 4 x 2 ) dy =*(14 y 2 ) dx; 

18. sin 2 y dx 4 2x cos 2y dy *» 0; 

20. dx ® (yx 3 — x) dy; 

22. (x 2 4 y 2 ) dx ~ xy dy; 

24. y 2 * (xy - xV)y'; 

0 . 

= p, the transformations 
dp 

=r » 

dx 

dp dy dp 


^ dx dy dx dy ^ 


(13-1) 


(13-2) 


often enable us to reduce an equation of second order in y to one of first 
order in p. For example, the equations 

F(x,y',y") = 0 , (13-3) 

F(y,y',y") = 0 (13-4) 

become by (13-1) and (13-2), respectively, 


44 ) -°- 

F (y,v,v 


dp\ 

dy) 


1 = 0 . 


(13-5) 


(13-6) 


These are first-order equations in p, and when p has been found, the sub- 
stitution p «■ y' yields a first-order equation for y. 



30 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

Example 1. Find the solution of y" sin y* ~ sin x which satisfies the conditions 
y(l) ** 2 and y'(l) * 1. 

Being free of y the equation has the form (13-3), and it becomes 

dp . 

sin p =* sin x 
dx 

by (13*1) . Solving this separable equation yields 

— cos p ** — cos x -f- c, 

which is reduced to p ** x by the condition p *» 1 at x » 1 . Writing dy/dx for p in the 
equation p ~ x gives on integration the final answer 

V = bzir 2 -l)+2 


in view of the condition ?/(l) *= 2 

Example 2. Solve yy" — 2 (y f ) 2 4 - y 1 =0. 

This equation has the form (KM), since it does not contain x. The transformation 
(13-2) gives 

yp - ~p~ + .i/ 2 = 0, 

dy 

which is a homogeneous equation with y as independent variable. Setting p *» vy and 
proceeding as in Sec. 7 give, after calculation, 

V - ±yV lTTV. (13-7) 

With p ** dy/dx in (13-7) we separate variables to obtain the final answers, 

x 4- cj — =F sinb"' 1 ~ • 

cy 


PROBLEMS 

Problems I and 2 are to be solved by the method of Example 1, Frobs 3 and 4 by that 
of Example 2, and Probs. 5 to 7 by whichever method is more suitable. 

1. (1 - x 2 )y" ~ xy f . 2. x(y" + y *) - y'. 

8. y" 4- * 0. 4. y" « yy'. 

8. x 2 y" - 1 - X. 6. yy" - y' 2 . 

7. xy" •* 4x - 2y'. 

8 . Solve y" ® 1 4- j; 7 * by both methods of this section, and verify the agreement of 
the results. 


GEOMETRY AND THE FIRST-ORDER EQUATION 

14, Orthogonal Trajectories. In a variety of practical investigations, 
it is desirable to determine the equation of a family of curves that intersect 
the curves of a given family at right angles. For example, it is known that 
the lines of equal potential, due to a distribution of steady current flowing 



GEOMETRY AND THE FIRST-ORDER EQUATION 


31 


SEC. 14] 

in a homogeneous conducting medium, intersect the lines of current flow 
at right angles. Again, the streamlines of a steady flow of liquid intersect 
the lines of equal velocity potential (see Chap. 7, Sec. 19) at right angles. 
Let the equation of the given family of curves be 

f(x f y,c) « 0, (14-1) 


where c is an arbitrary parameter. By specifying the values of the param- 
eter c, one obtains a family of curves (see solid curves in Fig. 6). Let 
it be required to determine the equa- 
tion of a family of curves orthogonal 
to the family defined by (14-1). 

The differential equation of the 
family of curves (14-1) can be obtained 
by eliminating the parameter c from 
(14-1) and its derivative, 


dx dy dx 


(14-2) 


Let the resulting differential equation be 

'WD- 0 - 



Now, by definition, the orthogonal family of curves cuts the curves of 
the given family (1 1-1) at right angles. Hence, the slope at any point 
of a curve of the orthogonal family is the negative reciprocal of the slope 
of the curves of the given family. Thus, the differential equation of the 
desired family of curves is 


F 



This is a differential equation of the first order, and its general solution 
has the form 

= 0. (14-3) 

The family of curves defined by (14-3) is the desired family of curves 
orthogonal to the curves of the given family (14-1). It is called the family 
of orthogonal trajectories . 

If the equation of a family of curves is given in polar coordinates as 
f(rfi,c) * 0, the tangent of the angle a made by the radius vector and the 
tangent line at any point (r,0) of a curve of the family is equal to r dB/dr 



32 


ORDINARY DIFFERENTIAL EQUATIONS 


[chap. 1 

(Fig. 7). Hence, by the preceding discussion, the differential equation 
of the orthogonal trajectories of the given family of curves is obtained by 
replacing rd$/dr by —dr/(rd$) in the differential equation of the given 
family of curves. 



Excumpk Let it be required to find the family of curves orthogonal to the family of 
curies (tig. Hj 


x 2 4 y 2 — cjr = 0 


( 1 4-4) 


The differential equation of the family (14-4) can l)e obtained by differentiating (14-4) 

with lespeet to x and eliminating the parameter 
c between (14-1) and the equation that results 
from the diffcientiation 

The reader w ill check that the differential equa- 
tion of the family (1 1-4) is 



dy 

2xv ' 4 x 2 
dx 


V 


0 . 


Hence, the differential equation of the family of 
curves orthogonal to (14-4) is 

dx 


2 xy 


dy 


x 2 4 y 2 “ 0. 


This is a homogeneous differential equation whose 
solution is found to be 

x 2 4 y 2 — cy *= 0. 


Thus, the desired family of curves is a family of circles with centers on the y axis (see 
Fig. 8). 



BBC. 15] 


GEOMETRY AND THE FIRST-ORDER EQUATION 


33 


PROBLEMS 


Sketch the following families of curves, find the orthogonal trajectories, and add them 
to your sketch: 




1. X* + y 2 - o J ; 

2. xy 

3. y « cx n ; 

. ** 


5 . r ** c; 

6. r » 

7. r m c( 1 — cos#); 

8. r * 


-cfl. 


V 


I — e cos 9 

9. If a and b are constant and X a parameter, show that the family of curves 

a* + X + 6 J + X 

satisfies an equation, free of X, which is unaltered when y f is replaced by — 1 /y'.* What 
does this indicate concerning the orthogonal trajectories? 

10. Fmd the algebraic equation, the differential equation, and the orthogonal trajec- 
tories for the family of circles tangent to the y axis at the origin. Verify your result by 
plane geometiy. (The configuration is a special case of so-called bipolar coordinates.) 

15. Parabolic Mirror. Pursuit Curves. Besides the problem of finding 
orthogonal trajectories, many other questions in geometry lead to first- 
order differential equations. The following examples show how geo- 
metrical conditions of this sort stem from physical conditions. Ine first 
is taken from optics, the second from 
the theory of pursuit. 

Example 1. Find a mirror such that 
light from a point source at the origin 0 is 
reflected in a beam parallel to the x axis. 

Let the ray of light. OP strike the mirror 
at P and be reflected along PR (Fig. 9). 

If PQ is the tangent at P arid or, 0, <£, and 
0 are the angles indicated, we have a » 0 
by the optical law of reflection and a * <t> 
by geometry. Hence 0 *» <f>. The equa- 
tion 



2 tan <f> 

1 — tan 2 4> 


gives 


tan 0 « tan (0 4- <j>) « tan 2 <t> 

v ^ 

x 1 - (V) 2 * 

since y ' « tan 4>. Solution of this quadratic equation for y ' gives 

, ~ X dc Vr 2 •+* y 2 


y 

xdx 4- y dy 
dt Vx 2 4- ^ 


dx. 


whence 



34 ORDINARY DIFFERENTIAL EQUATIONS {CHAP. 1 

The left-hand member of this is an exact differential, and we get, on integrating, 

±v?T7 «* x 4* c, 

which, on squaring, yields y 2 «* 2cj -f c 2 . The curves form a family of parabolas with 
focus at the origin. 

Bxample 2. A boat A moves along the y axis with constant speed a. Find the path of 
a second boat B which moves in the left-hand half of the xy plane with constant speed 
b and always points directly at A . 


At a time t inin after A is at (0,0), we shall have A at (0,ctf) and B at (x,y), say. Since 
the line AB is tangent to the path of B, the slope of this line equals the slope of the path, 
so that 


y — at dy 
x — 0 dx 



or xy' y *=> —at. 

To eliminate t , we first differentiate (15-1) and obtain 


(15-1) 

C xy ' - vY - xy” * -o~* 
dx 


(15-2) 

Since ds/dt « b, where « is an arc on the trajectory, we have 


dt dt ds 1 / -r 

dx ds dx b T y 


(15-3) 

With r defined as a/6, substituting (15-3) in (15-2) yields 



xy" « — rVT "+V 5 , t - p 

b 


(15-4) 

which is reduced to a separable equation of first order by letting p *» y' 
The solution is 

,' = P » a inh(rloK-) --[(-)-(-)] 

as in Sec. 13. 

(15-5) 


and from this, y is found by integration. 


PROBLEMS 

Find the curves in the xy plane which satisfy the following conditions: 

1. (a) The tangents pass through the origin; (6) the normals pass through the origin. 

2. (a) The segment of tangent between a point on the curve and the x axis has unit 
length; (6) the projection on the x axis of this segment has unit length. 

3. (o) The area bounded by the curve, the x axis, and the ordinate equals the ordinate; 
(6) the area equals the length of the curve from (0,1) to (x,y). 

4 . Find the path of a small boat in a wide river with uniform current if the boat has 
constant speed relative to the water and always heads toward a fixed point on the bank. 

6. Solve Example 2 completely under the assumption that A is at (0,0) and B is at 
(*o,0), at time f « 0. Distinguish the cases r «■ 1 and r & 1. If r < 1, at what point 
and when does B overtake A? If r « 1, how close can B get to A ? 

16 * Singular Solutions. It was remarked in Sec. I that a differential 
equation may possess singular solutions, that is, solutions which cannot 



SEC. 16] GEOMETRY AND THE FIRST-ORDER EQUATION 35 

be obtained from the general solution by specifying the arbitrary constants. 
For investigation of this phenomenon let the family of integral curves 
defined by 

4 >(*,y>c) ~ 0 (16-1) 

be the general solution of the first-order equation 

F(z,y>y') * o. (16-2) 

Assume that the family of curves (16-1) possesses an envelope, that is, a 
fixed curve C such that every member of the family is tangent to C and 
such that C is tangent, at each of its points, to some member of the family. 
At a point (x,y) on the envelope, the values x , y , y ' for the envelope are 
the same as for the integral curve, and hence these values x, y , y' satisfy 
(16-2). Thus an envelope of a family of solutions is again a solution. 

In general, the envelope is not a curve belonging to the family of curves 
defined by (16-1), and hence its equation cannot be obtained from (16-1) 
by specifying the value of the arbitrary constant c. It is known from cal- 
culus that the equation of the envelope is obtained by eliminating the 
parameter c between the equations 

<t>(x,y,c) « 0 and 4> c (x,y,c) = 0, 

where <f> c &* d<t>/dc. 


Example: The family of integral curves associated with the equation 


is the family of circles 


y 2 (y') 2 + v 2 » 

(x - c) 2 + y 1 - a 2 . 


(16-3) 

(16-4) 


The equation of the envelope of the family (16-4) is obtained by eliminating c between 
(16-4) and <t> e ■* ~2{x — c) «* 0. There results 


V - zka, (16-5) 


which represents the equation of a pair of 
lines tangent to the family of circles (16-4) 
(Fig. 10). Obviously, (16-6) is a singular 
solution of (16-3), for it is a solution, and 
it cannot be obtained from (16-4) by any 
choice of the constant c. On referring to 
Sec. 1, it is easy to check that the condi- 
tions of the theorem ensuring uniqueness 
of the solution are violated in this example. 



PROBLEMS 

1. (a) Show that y — c *» (as — c) ! represents a family of congruent parabolas with 
vertex on the line y *■ x, and sketch. (6) By differentiating with respect to c obtain 
the envelope y « % — (c) By direct computation, verify that the parabolae and the 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 

envelope have the same slope at corresponding points, (d) Obtain a first-order differen- 
tial equation for the family, and (<?) verify tirnt y « x — M is a singular solution of this 
equation. 

2. A particle on the z axis has velocity v = \^s, where 8 is the distance to the origin. 
Show that the motion is uniquely determined if the particle is at any point other than 
the origin but that infinitely many different behaviors can occur if the particle ever 
reaches the origin. 

S. (a) Obtain the equation yy' + y' 2 -f x » 0 for the orthogonal trajectories of the 
family y «* cx + 1/r. (b) Show that y » 2 \/z is the envelope of the family. ( c ) At 
points of the curve y ** 2 Vs find the slope of the solutions of yy' + ( y f ) 2 + x « 0 in 
terms of z. Then find the slope of the curve y * 2-\/x in terms of x How are these 
two slopes related? Why? (d) Sketch the family, the envelope, and the orthogonal 
trajectories in a single diagram. 

17. The General Behavior of Solutions. The foregoing paragraphs in- 
dicate* that from suitable geometric conditions on a curve, one can obtain 
a differential equation for the curve. Now, in this section the point of 
view is to be reversed. Starting from the differential equation we obtain 
certain geometric conditions, which enable us to describe the solution 
qualitatively even when the equation itself cannot be solved. 

The function f(x 9 y) in the general first-order equation 

dy 

7 - (17-1) 

ax 

gives the slope of the solution curve at each point (x,y). Hence the solution 
curves are increasing functions of x in regions of the xy plane in which 
/(#,#) is positive and decreasing in regions where f(x,y) is negative. For 
continuous/^,?/) the boundary between these regions is part or all of the 
curve 

f(*>y) * 0. (17-2) 

Equation (17-2) gives the locus of the critical points, and their character 

(maximum, minimum, neither) is shown by the sign of f(x,y) at neighbor- 
ing points. The inflection points and sense of concavity are similarly 
found from 

y" -U+Uv’ =/*+/*/, (17-3) 

where f x ss df/dx and f v ss df/dy. 

For more detailed information one can plot the curves 

/(*,!/) * c, (17-4) 

called isoclines . At any point (x } y) where (17-4) holds, the solution curve 
approximates a straight-line segment of slope c, a fact which can be used 
as a check on the qualitative information obtained from (17-2) and (17-3). 
From this viewpoint (17-1) is equivalent to a direction field in the xy 
plane as discussed in Sec. 1. Any curve whose tangent at each point has 



37 


SB€. 17] GEOMETRY AND THE FIRST-ORDER EQUATION 

the direction of the field is a solution, and conversely. The isoclines 
(17-4) and the direction field discussed in Sec. 1 lie at the basis of some 
methods for numerical solution of differential equations. 



A technique of obtaining approximate solutions, based on a comparison 
idea, was developed bv S. A. Chaplygin for equations of first and higher 
orders (see Fig. 11). Let y\(x), y(x), and 7/2 M be solutions of 

dy 1 dy dyo 

7 = /(*,»), -p -/«(*,» 2). (17-5) 

ax ax dx 

By subtraction, the difference y — y\ satisfies 
d 

— (y - yi) * K*,y) - v = y(*), 2/1 * yi(*). (17-6) 

ax 


Now, if /(x,?y) > f\(x,!h) in a range of x, then y — is an increasing 
function of x in that range. In this case the condition y ~~ y x = 0 at some 
point x 0 ensures that ?y — */i > 0 for x > x 0 and y — y\ < 0 for x < x 0 . 
Similar remarks apply to y 2 — Hence the conditions 

> f(x,y) > h(x,y), 

(17-7) 

2/1 (Xo) = 2 /(*o) = 3/2(10), 


in (17-5) enable us to conclude that 

2/i (x) > y(x) > y 2 (x), 

Vi(x) < y(x) < y 2 {x), 

One chooses /i (x,y) and f 2 (x,y) in such a way that the solutions y u y 2 are 
obtainable by elementary methods. Equation (17-8) then gives an explicit 
estimate for y(x). 


x > x 0l 
x < x 0 . 


(17-8) 


A refinement of these ideas leads to an explicit and important inequality for estimat- 
ing the error in certain approximations. Let y(x) be an exact solution of y' — f(x,y) 



ORDINARY DIFFERENTIAL EQUATIONS 


38 


[chap. 1 


through the point (xo,y 0 ), and let y\(x) be an approximate solution through this point. 
Substituting y x {x) into the equation gives 


rg-fanMl + H*}. 


(17-9) 


where the error term e(x) arises because yi is not an exact solution. Now, what can be 
said about the solution error | ij\ ~ y\ in terms of the substitution error e(x)? 

To answer this question we suppose that f(x,y) is continuous in a region containing 
(xo,y 0 ) and satisfies a so-called bipschitz condition there; that is, 

\f(x,y{) -f(x t y)\ < k\yi - y\, k const, (17-10) 

for some k and all x, y, and y\ in the region. The condition (17-10) stipulates that 
f(x,y) shall not change too rapidly when y changes. In case f y exists, the mean-value 
theorem gives 

f(*>y) -f(x,vi) - fv(x,Z)(y - vi), y <k <yu (17-11) 

and hence (17-10) holds in any region throughout which 

\fv(x,y)\ <k. (17-12) 

Returning to the original question, in (17-9) let E( x) be the error in y Xi 

E(x) » yi (x) - y(x). (17-13) 

Since y(x) is an exact solution, we have y f » f(x,y ), and hen<*e, subtracting from (17-9), 
dih dy dE 

^ - f(x, Vl ) - f(x,y) + e(x). (17-14) 

dx ax dx 


If (17-10) holds, and if | e(r) | < m, then (17-14) leads to 


dE 

dx 


< \f(x,yi) - f(x,y) 1 + | e(x) | 


< k\y\ — y\ + m *= k\ E(x) | -f m. (17-15) 


If we could drop the absolute values in (17-15) and replace the < by 
tain the linear equation 

~ ** kE{x) + m. 
dx 

The solution with E(xq) ** 0 is 


E(x) - ~ (f*' 1 -*. 1 - 1). 
k 


we should ob- 
(17-10) 

(17-17) 


Now, it is plausible and can be proved rigorously that E(x) in (17-17) is the maximum 
possible E(x) subject to (17-15), with x > xq. Hence the solution error E *» yi — y 
satisfies 


|»i(x) -y(x)\ ^(e*l’-*„l - 1), 


{ m *» max | e(x ) |, 
k «* Lipschitz constant, 


(17-18) 


where \x — xo| is used rather than (x — xo) to account for the case x < x 0 . 

Equation (17-18) leads at once to a uniqueness theorem, for if y\(x) is an exact solu- 
tion, then e(x) * 0 in (17-9), hence m * 0 in (17-18), and therefore ydx) « y(x). 

Example 1 . Discuss the integral curves for the equation y' ** xy — 1 without solving 
the equation. 



GEOMETRY AND THE FIRST-ORDER EQUATION 


SEC, 17] 


39 


The hyperbola xy « 1 is the locus where y' ** 0. If xy > 1, then y* > 0 and the 
solution curves are increasing, but if xy < 1, they are decreasing. Hence, xy » 1 gives 
a locus of minima in the first quadrant, maxima in the third quadrant. Since y f ** — 1 
when x « 0 or y * 0, all integral curves intersect the axes at an angle of 135°. From 

V" « xy' -{- y * x(xy - 1) -f y * y(x 2 -f 1) - x, 

the curve is concave up if y > x/(x 2 1) and concave down when this inequality is 

reversed. The curves have the appearance shown in Fig. 12. 



Fio. 12 


Example 2. If r/' *» sin xy , j/(0) ** 1, show that 

e x * /r < y < e* t/2 , 

at least for 0 < x < 0.8. 

A glance at the graph of sin u shows that 

2 

- u < sin u < u 
x 


for 0 < u < t/ 2, and hence 


2 



40 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

The solution of the given equatio therefore lies between the solutions to 

, 2 

y ** - xy, y « xy, 

r 

This gives the desired inequality for the range 0 < xy <; ir/2 . Since y < e*^ 2 , it suffices 
to have 

0 < xe** 12 < ir/2, 

and this is true for 0 < x < 0.8. 

PROBLEMS^ 

1. For the equation y ' * y/x — 1, (a) sketohlthe locus y' ** 0 in the xy plane, (b) 
Indicate the regions in which y is increasing; decreasing, (r) When is y concave up? 
Down? ( d ) At what slope do the solutions cross the axes? (e) Sketch the locus where 
the solutions have slope 1, —1,2, —2, 5, —5. (/) Sketch the solutions as well as you 
can. (g) Verify your work by solving the equation. 

2. In what regions of the xy plane are the solutions of y' = sin ( x 2 + y z ) increasing? 
Decreasing? Sketch. 

3. Discuss the equation y ' ~ sin (x 2 -f y), y{ 0) * 2, by comparing with suitably chosen 
simpler equations. 

APPLICATIONS OF FIRST-ORDER EQUATIONS 

18. The Hanging Chain. Let it be required to find the curve assumed 
by a flexible chain in equilibrium under gravity (Fig. 13). With $ as arc 


dx 


Fig. 13 

from the point x = 0, let the weight density of the chain be w($) lb per ft 
and let the loading function be f(x) lb per ft. The equation of the curve 




SBC. 18] APPLICATIONS OP FIRST-ORDER EQUATIONS 41 

y y(x) will be obtained from the fact that the portion of chain between 
0 and x is in static equilibrium. 

Equating horizontal forces gives 

T 0 - T cos 8 , (18-1) 

where To is the tension at the lowest point and 8 the angle which the tan- 
gent to the curve makes with the horizontal. Similarly, equating vertical 
forces gives 

w(s) ds + J Q f( x ) dx ~ T sin 8 (18-2) 

since the weight of the chain-plus-load must be balanced by the vertical 
component of T. Both (18-1) and (18-2) require that the function y{x) 
be differentiable (so that 8 is well defined) and they use the fact that the 
tension is tangential, for a flexible chain. 

From (18-1) we have T - T 0 /cos ft, so that 

T sin 8 = To tan 6 « Toy', (18-3) 

the latter equation resulting from the definition of ?/ as slope. Substitution 
in (18-2) gives 

f w(s) ds + f f(x) dx = T 0 y'. (18-4) 

J o Jq 

When w and / are continuous, (18-4) may be differentiated with respect 
to x, a procedure which leads to the differential equation for the curve 


in view of the fact that 


ds 

w(s) — + f(x) 
dx 


Toy " 


d r* 

* — / w(s) ds 

dx J 0 



w(s) ds 


ds 

dx 


(18-5) 


Example Show that a uniform flexible chain acted upon by gravitational forces alone 
assumes the shape of a catenary, and find the tension in terms of the height y . 

Here f(x) * 0, w($) ® m, a constant. Since d*/dx » Vl + y 7 \ Eq. (18-5) gives 

woVTTV* - T«y". (18-6) 


This is a second-order equation, which can be reduced to one ot first order by the method 
of Sec. 13. With p » dy/dx we have 


and hence (18-6) becomes 


dp dp dy 

— - sec — — 

dx dy dx 


cV X 4 ~ P 2 




(1M) 


hnS' q 


4 



42 ORDINARY DIFFERENTIAL EQUATIONS 

Hub equation is separable, the solution being 


[chap. 1 


c y - vTT? (18-8) 

when the axis is so chosen that cy » 1 when p ® 0. Equations (18-1) and (18-8) now 
give T » To sec 9 - To V 1 -f tan 2 0 * T 0 V 1 -f p 2 « Toq/ * tooy, which gives the 
tension. Writing p dy/cte in (18-8) and separating variables yield 


as the reader will verify. 


To . u'qx 
— cosh — 
t^o T 0 


PROBLEMS 


1. A flexible weightless cable supports a uniform roadway weighing m lb per ft. 
The tensions at the highest and lowest points are T and To, the roadway is 2a ft long, the 
sag is b, and the length of the cable is 2s. If the cable is symmetric about the y axis 
and has its lowest point at the origin, show that the equation of the curve is 

u>oz 2 

y at , 

2 T 0 

and thus obtain the relations T 0 * woo 1 /2b, T *» u\){afb)\/ g 2 /4 -f b 2 , 

8 ~ f Vl -f (2bx/a?)* dx, a — a ~ 2b 2 /3a 
Jo 

as b 0 . Hint: \/ 1 -f u ~ 1 ~ m /2 as u — ♦ 0 . 

2. One end of a flexible uniform telephone wire is b ft above the lowest point, at a 
distance a ft from it measured horizontally and at a distance $ ft from it along the wire. 
If u =® awo/To, in the notation of See. 18, show that u satisfies the transcendental equa- 
tions (cosh u — l)/u ** b/a, (sinhu)/u ** a/a and hence by division the nontrans- 
cendental vquation tanh (m/2) » b/a. Also find the relations To « woa/u ** wqs each m, 
T — wqCi (cosh u) / u « wqs coth u for Tq and for T, the tension at the highest point. 
The student familiar with infinite series will obtain simplified expressions by expansion 
of the hyperbolic functions when u is small, that is, when the tension is large. 

19. Newton’s Law of Motion. Newton’s second law' of motion states that 
the time rate of change of momentum is equal to the impressed force. In 
symbols, 

— (™*>) = F, ( 19 - 1 ) 

at 

where F = component of force in the direction of motion 
m * mass of the moving particle 
v *» ds/dt = velocity of the moving particle 
It is supposed that the particle moves in a straight line, its distance from 
some fixed point on that line being 8. 

The differential equation (19-1) is quite general, since the force may 
depend on the time t } on the displacement s T and, in the case of damped 
motion, on the velocity v . Also the mass may be variable in some problems, 
for example, those concerned with rocket flight or with high-speed electrons. 



SBC. 19] 
Since 


APPLICATION S OP FIRST-ORDER EQUATIONS 
d(mv) d(mv) ds d(mv) 

S* a* V 

dt ds dt ds 


Eq. (19-1) may be put in the form 

d(mv) 

v 

ds 


F, 


43 

(19-2) 


(19-3) 


which gives an alternative statement of Newton's law. Multiplying 
(19-3) by m leads to ( mv)d(mv)/ds = Fm, which may be written 

d 1 

*7 ” (mv) 2 » Fm. (19-4) 

CIS A 

This gives still another formulation. 

In case m and F are known in terras of s only, m * m($), F » F(s), then 
(19-4) may be solved completely: 

Vzbrw) 2 - J/ 2 (mot>o) 2 = f F(s)m(s) ds. (19-5) 

•'•O 


If the mass is constant, (19-5) becomes 


Yimv 2 — }imv § » / F(s) ds, (19-6) 

since m(s) may be factored out of the integral. Then (19-6) is the law of 
conservation of energy , for the left side of (19-6) is the change in kinetic 
energy while the right side represents the work done when the particle 
moves from s 0 to s. Thus the right side is the change in potential energy. 
The steps leading to (19-6) are evidently reversible if F(s) is continuous, 
and hence Ne wton's law is equivalent to the principle of conservation of energy, 
when the mass is constant and the force ts a continuous function of position 
only . 

When F and m are known functions of 8 , it has been seen that one can 
obtain a so-called first integral of the equations. If F is a known function 
of l, the same is true; we have 

mv — m 0 v o = f F(t) dt, (19-7) 

Jto 

by inspection of (19-1). And similarly, when F and m are known functions 
of v, one can write d(mv) = m(v) dv + vm'(v) dv. Substitution in (19-1) and 
separation of variables now give 


m(v) + vm f (v) 



[chap. X 


44 ORDINARY DIFFERENTIAL EQUATIONS 

The same process in (19-3) yields 


8 — S 0 



m(v) + vm'(v) 
F(v) 


vdv . 


For several particles addition gives 



XFi 


(19-9) 


(19-10) 


With M as the total mass M «* 2m, and with V as the mean velocity, MV * 
this may be written (d/dl)(MV) « F. Here F * 2 F % is the total force; but since the 
internal forces cancel in pairs, by Newton’s law of equal and opposite reaction, F is also 
the total external force acting on the system. The extension to continuous mass dis- 
tributions is made by analogy, the equations being defined as the limiting form of those 
for a set of approximating discrete distributions. Thus any point moving with the mean 
velocity V satisfies Newton’s law in the form (19-1). It can be shown that this point 
actually remains “inside” the body if the v t are suitably restricted, but some restriction 
is necessary. Of course, if the masses are constant, then V =* dS/dt, where S is the posi- 
tion of the center of mass, MS — 2m,s, In that case the center of mass itself follows 
<1M). 

Example 1. The force on a particle of mass m is proportional to its distance from the 
origin and is directed toward the origin. Find a differential equation for its motion. 

The force is ks if a is the distance from the origin at time t Since the force is directed 
toward the origin, it has at all times a sign opposite to that of s. Thus k is negative, 
and one may write k *» — mco 2 for some constant o>. Equation (19-1) will now give 
d(mv)/dt « -Tito 2 * or, dividing by m and putting v ** ds/dt, 

^j + «A-0. (19-11) 


This is the equation for simple harmonic motion, an important type of periodic motion 
that arises in many mechanical a nd electrical systems. The general solution of (19-11) is 

s =* A cm (uit -f- B) (19-12) 


as shown in Probs. 2 and 3 and in the Example of Bee. 21 . Hence the motion is periodic, 
with period 2ir/w independent of the amplitude A and plmse B. 

Example 2, A gun containing a bullet moves with nonnegative velocity v on a straight, 
horizontal, frictionless track and points in a direction exactly opposite to that of the 
motion. The mass of builet-plus-gun is m, and that of the bullet is —Am, where Am Is 
negative. If the bullet is fired with velocity r relative to the gun, show that (v — r) Am 
equals the momentum of the gun after firing minus the momentum of the bullet-plus- 
gun before firing. 

By (19-1) the momentum of the bullet-pl us-gun is constant, since there is (we as- 
sume) no external force on this system as a whole. Hence 


mv « (m -f Am)(t> -f Aw) 4* ( — Am)v&, (19-13) 

where p + At> is the new velocity of the gun and v b of the bullet: 

% - v - c. (19-14) 

Computing (m + Am)(t> -j- Av) — nw from (19-13) and (19-14) gives the result. 



SBC. 19] APPLICATIONS OP FIRST-ORDER EQUATIONS 45 


Two remarks are in order. First, the “equation of continuity for momentum/' Eq. 
(19-13), has been seen to follow from Newton’s law; it is not a new assumption* Second, 
if one replaces (19-14) by 

«6 * v -f Ai> - c (19-15) 


(which is a justifiable alternative), the result is altered only by the second-order term 
Av Am. Hence there is no change at all when the increments are replaced by differentials 
as in the following example. 

Example 3. A. rocket fires some of its mass backward at a constant rate r kg per sec 
and at a constant speed c m per sec relative to the rocket. Show that the thrust devel- 
oped is r(c — v) when the velocity of the rocket is v. If the rocket starts with velocity 
eo and there is no other force acting, v vq 4 c log 2 when half the mass is used up. 

With m and v the mass and velocity of the rocket at time t, the differential in momen- 
tum d(mv) due to external forces is F dt by (19-1). That due to loss of a mass —dm at 
speed c — v in the backward direction is d(rnv) « (c — v)( —dm) by Example 2; for 
differentials, not increments, the result is exact. Thus is obtained a fundamental rela- 
tion for rocket problems: 

d(ntv) - Fdi - (c - v) dm. (19-16) 

In the present case dm « —r dt, so that (19-16) gives 


d{mv) 

~~dT 


F *f r(c — v) 


(1 9-17) 


after division by dt. Hence the effect of the rocket motor is to add r(c — v) to the force F, 
and that is what was to be shown. 

Substituting 

m — mo - rt (19-18) 


for m in (19-17) gives (mo — rt)(dv/dt) « rc -f F after slight simplification. Hence, by 
separating variables, 

v — tty * ( r + - ) log for constant F. (19-19) 

\ r / m 


Putting F ® 0, m « mo/2 gives the second result. 

Example 4. Starting with velocity t>o an electron is accelerated for a distance s by a 
constant electric field of magnitude E. What is the terminal velocity? 

Let c be the velocity of light, so tha'. the mass m of the electron is given in terms of 
its rest mass mo and its velocity v by 

« - — r^W (19*20) 

v 1 - ir/r* 

If we write 


■ sin 8. 


V-l 


cos 0 


(19-21) 


as we may for v < c, then ( 19 - 20 ) gives m « m 0 sec 8 and mv ** cm tan 6. Substituting 
in (19-3) with F « Ee, where e is the charge on the electron, gives 

F eE 


de 


(rmo tan (?) 


- esc 8. 


Hence sec 2 8(dd/ds) - {eE/rtw?) esc 8, and by integration 


seE 

sec 8 • sec 0o + — « 
me 


(19-22) 



m 


ORDINARY DIFFERENTIAL EQUATIONS 


(CHAP. 1 


where 0o refers to the initial value. For numerical calculation it is more efficient to use 
the form (19-22) with trigonometric tables than to obtain v explicitly by (19-21). 


PROBLEMS 

1* A brick is set moving in a straight line over ice with an initial velocity of 20 fps. 
If the coefficient of friction between the brick and the ice is 0.2, how long will it be before 
the brick stops? 

2. Find a value of a for which 8 « ci sin (at -f- Oi) is a solution of (19-11). Does this 
expression have enough independent constants to be a general solution? Determine 
Ci and cj in such a way that the displacement s is maximum at t » 0 and has then the 
value A. Determine c\ and e? in such a way that the maximum displacement is A and 
the maximum velocity occurs at t ~ 0. 

3. Apply the transformations used in the derivation of (19-6) to obtain the appropri- 
ate form of (19-6) from (19-11). Check by direct comparison with (19-6). Solve the 
resulting equation by separating variables, and thus show that the solution obtained in 
Prob. 2 includes every solution. 

4 . Suppose the rocket in Example 3 is subject to a retarding force of magnitude 
mg *+■ fa;, where g and k are constant. From (19-17) and (19-18) obtain a linear equation 
for v as a function of t. Show that (mo — rt)~ klr is an integrating factor, solve for 

«, and obtain the position s at time t from s ~ Jvdt. 

6. The equation of a cycloid Is x «* 8 + sin 6, y * 1 — cos 0. Show that the arc « 
from the lowest point satisfies $ 2 = 8 y t and deduce the equation 4c 2 = gr(«o — s 2 ) for 
a particle sliding down the curve. By differentiation obtain the equation 4 d 2 s/dt 2 * 
— g* t which shows that the motion is simple harmonic. What is the period? 

6. In a microwave electron accelerator the field is E sin u>t, If an electron starts with 
velocity t*>, find the maximum possible terminal velocity. Hint: The maximum occurs 
when the time for passage is exactly *-/<»>, for an electron starting at time t «* 0. Use 
(19-8). 

20. Newton’s Law of Gravitation. Another law of Newton is the law of 
gravitation, to which he was led in his attempt to explain the motion of 
the planets. This law states that two bodies attract each other with a force 
proportional to the product of their masses and inversely proportional to 
the square of the distance between them, the distance being large compared 
with the dimensions of the bodies. If the force of attraction is denoted 
by F, the masses of the two bodies by m x and m 2 , and the distance between 
them by r, then 

ymim 2 


where y is a proportionality constant, called the gravitational constant. 
In the cgs system the value of y is 6.664 X 10~ 8 . 

It can be established that a uniform spherical shell attracts a particle 
at an external point as if the whole mass of the shell were collected at the 
center (see Chap. 5, Sec. 14). Hence, by integration, the same is true for 
a solid sphere provided the density is a function of the radius only. If the 



47 


SEC. 20] APPLICATIONS OP FIHST-OKDEB EQUATIONS 

sphere is the earth, one can therefore write 

F = m v(~) toward the earth, (20-2) 

where r e * earth’s mean radius 

r « distance from particle to center of earth 
g = new constant called acceleration of gravity 
Its value in the cgs system is approximately 980 cm per Bee per sec and 
in the fps system 32.2 ft per sec per sec. Since the earth is not a perfect 
sphere, and since the density varies from place to place, the value of g 
depends slightly on location. One uses a plus or a minus sign in (20-2) 
according as the positive direction is taken toward or away from the earth’s 
center. 

It can be shown that a uniform spherical shell exerts no force on a particle which is in 
the hollow space enclosed by the shell (Chap. 5, Sec. 14). Hence the force on a particle 
of mass m at distance r from the center of a sphere is 

™ j 4ttu 2 p(u) du (20*3) 

when the density of the material forming the sphere is a continuous function p(u) of the 
distance u to the center The special case p(u) = 0 for u > r 0 gives the result for a par- 
ticle outside the sphere, as discussed previously. 

Equation (20-3) gives 

F » mg ~ (20-4) 

for a particle inside the earth if p is taken as constant In case the particle is close to 
the surface, we have r ~ r*, so that either (20-2) or (20-4) takes the simple form 

F * mg. (20-5) 

The error in (20-5) is less than 1 per cent for heights up to about 20 miles. 

Example 1. Neglecting air resistance, discuss the velocity of a particle falling toward 
the earth. 

The principle of conservation of energy combines with (20-2) to give 



or, after carrying out the integration, 

( 20 - 6 ) 

If the particle starts from rest at a very great distance, the velocity with which it strikes 
the earth is 

v e *» \^2gr e (20*7) 

as we see by setting Vo « 0, ro « *>, r « r e in (20-6). This terminal velocity is also the 
minimum velocity of escape for a particle which leaves the earth never to return. Since 
is approximately 4,000 miles, we find from (20-7) that v e is nearly 7 miles per sec. 



48 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 


Example 2. Obtain a differential equation relating the density and pressure in the 
interior of a spherical star if each is a function of the distance to the center only: /> « p(r), 
p ** p(r). Assume p(r) continuous. 

Consider a column of material of unit cross section extending along a radius from 
r to r, t the radius of the star. The pressure at the base of this column equals the total 
downward force on the column due to gravitation. The differential force on an element 
of the column from r m q to r * q + dq is given by 


df - 7 — ^ £ jTVuVu) du J a <*>(?) dq, 


in accordance with (20-3), and the total force is given by integration: 


p(r) « F * f 9 4>(<j) dq f <*>(<?) dq . 

Jr Jrs 

Since the right side of Eq. (20-8) has a continuous derivative, so does p{r): 

= - 7 —? [ 4tt u 2 p(u)du. 
dr r y 0 

Multiplying by r 2 /p(r) and differentiating again lead to 

l(;S) + ' Wr2p ” 0 ' 


(20-8) 


Example 3, Assuming conservation of energy for motion in a curved path, obtain an 
expression giving the period of a simple pendulum. 



Fig. 14 


Let P denote the position of a pendulum bob suspended from 0, and let 6 be the 
angle made by OP with the position of equilibrium OQ t as shown in Fig. 14. The work 
required to change 6 to any other value a is the work required to raise the bob through a 
vertical distance a cos B — a cos <*, if a is the pendulum length. With a chosen as the 
angle for maximum displacement, so that v ■* 0 at 6 as, conservation of energy gives 


1 

2 



2™ 


® » mga(coB B — cos a), 



SEC. 20] APPLICATIONS OP FIRST-ORDER EQUATIONS 49 

where m is the mass of the bob. Separating variables gives t in terms of $ and hence, 
implicitly, 8 in terms of U In particular the time required for $ to increase from 0 to a is 


T 

4 


%/lf: 


ds 


Zg J o vco s 0 — cos a 


(20-9) 


so that the period T depends on the amplitude a. This dependence leads to the so-called 
circular error in pendulum clocks. 

The identities cos 0 «* 1 — 2 sin 2 (0/2), cos a * 1 — 2 sin 2 (a/2) give 

fa r do a 

T - 2*/- / k m sin-- (20-10) 

V g Jo Vk 2 - sm 2 (0/2) 2 

If a new variable of integration 4> is defined by 



k sin 4>, 


( 20 - 11 ) 


them ^ ranges From 0 to r/2 when 6 ranges from 0 to a. Also by (20-11) 
2k cos 4> d<t> 2VT 2 — sin 2 (0/2) 

(IS jzszrzz-r- rr- ^rrr - CW>. 

cos 0/2 \/l — fc 2 sin 2 </> 

Substitution into (20-10) yields a so-called elliptic integral 



d(f> 

\/ 1 ~ k 2 sin' 2 (f> 


k 


a 

mma ? 


( 20 - 12 ) 


The advantage of (20-12) over (20-9) is that (20-12) has been extensively studied and is 
available in tables A series expansion is easily obtained, by expanding the radical 
for small a. The result is 


r. + (I)V + QV + (11|)>,...]. 


The function 


F(k,x) « f ~ 
Jo ^ 


d<t> 


\/ 1 — k 2 sin 2 0 

is called the elliptic integral of the first kind. See Chap. 2, Sec. 10. 


PROBLEMS 

1, A stone is thrown vertically upward with velocity 8 fps at time t 0. Using 
(20-5), write an expression for the position and velocity at time t and also for the velocity 
as a function of distance 8. Find the time at which the velocity is zero, and show that 
the height is then maximum. Show' that the maximum height agrees with that obtained 
by equating kinetic and potential energy, that is, with nigh ** mv o/2. 

2 . A particle slides down an inclined plane, making an angle 0 with the horizontal. 
If the initial velocity is zero and friction may be neglected, the component of force in 
the direction of motion is F * mg sin 0, What are the velocity of the particle and the 
distance traveled during the time f? Find the speed as a function of the vertical distance 
fallen, and verify that the same result would be given by equating energies as in Prob. 1. 



SO ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

At &ny given instant, show that the locus of the particles obtained for various fts is a 
circle (Fig. 15). 

3. Suppose the pressure in the atmosphere is a known function of the density, p » f{p). 
Show that the height h at which the density has first dropped to zero from the sea-level 

r*> 

value po satisfies gh/{\ -J~ h/r 0 ) m j [/'( p )/ p ! dp, when (20-2) is used. Hence the de- 
pendence of gravitation on distance introduces an effect which depends on h only, not 
on/(p), and the effect is less than 0.02/i when h is less than 80 miles. Obtain an explicit 
expression for h in the case of adiabatic expansion, p ** kp a . For what values of a does 
this give a finite height? (For air a » 1 .5; the height thus obtained turns out to be about 
18 miles.) 



4 . A man and a parachute weighing w lb fall from rest under the force of gravity. 
If the air resistance is proportional to the square of the speed v, and if the limiting speed 
is vo, find the speed as a function of the time ( and as a function of the distance fallen s 
Hint * (w/g)(dv/dt) — w — kv 2 . 

5 . A projectile is fired, with an initial velocity vq, at an angle a with the horizontal 
Find the equation of the path under the assumption that the force of gravity is the only 
force acting ori the projectile. For what a is the range maximum? Describe the region 
which is within the range of the gun. Hint: Find the envelope of the trajectories when 
vo is fixed but a varies. 

6. A cylindrical tumbler containing liquid is rotated with a constant angular velocity 
about the axis of the tumbler. Show that the surface of the liquid assumes the shape of 
a paraboloid of revolution. Hint: The resultant force acting on a particle of the liquid is 
directed normally to the surface. This resultant is compounded of the force of gravity 
and the centrifugal force, since pressure at the free boundary is zero. 

7 . Water is flowing out through a circular hole in the side of a cylindrical tank 2 ft 
in diameter. The velocity of the water in the jet has the value which it would attain 
by a free fall through a distance equal to the head. How long will it take the water to 
fall from a height of 25 ft to a height of 9 ft above the orifice if the stream of water is 
1 in. in diameter? 

8 . Water is flowing out from a 2-in. horizontal pipe running full. Find the discharge 
in cubic feet per second if the jet of water strikes the ground 4 ft beyond the end of the 
pipe when the pipe is 2 ft above the ground. 



SBC* 21] 


LINEAR DIFFERENTIAL EQUATIONS 


51 


UNBAR DIFFERENTIAL EQUATIONS 

21. Linear Homogeneous Second-order Equations. An equation of the 
form 

V" + Pi(*)v' + V2 &)y =* 0, (21-1) 

in which p\(x) and p 2 (x) are specified continuous functions of x in a given 
interval ( a,b ), is called a linear homogeneous equation of second order. 
From the existence and uniqueness theorem of Sec. 1 it follows that this 
equation has a unique solution for every x = x 0 in (a, b), satisfying the 
initial conditions y(x 0 ) « yo> 2/'(x 0 ) ~ Vo- Thus, the integral curve for 
Eq. (21-1) is determined uniquely when the ordinate and the slope of the 
curve are specified at a given point of the interval. 

Equation (21-1) is called linear because its solutions satisfy the following 
linearity properties: 

1. If y yi(x) is a solution of (21-1), then y = cy i(x), where c is a 
constant, is also a solution. 

2. I f y — i/i (x) and y — y 2 (x) are two solutions of (21-1), then their 
sum y ~ i/i (x) + y 2 (x) is also a solution. 

It follows from these properties that the sum of any number of solutions 
of (21-1) each multiplied by a constant is also a solution. 

The proof that properties 1 and 2 hold is simple 

Thus, suppose that y = yi(x) is a solution of (21-1); then the substitu- 
tion in (21-1) gives an identity 

y[ + Pi y'l + VzVi 32 o. (21-2) 

We must show that 

fa/i)" + pi (cy)' + P 2 ^yi) (21-3) 

vanishes identically for every constant c. But since c can be taken outside 
the differentiation sign, we can write (21-3) as 

c(y[ + pm + p 2 yi) } 

and this vanishes because (21-2) does. 

This establishes property 1. 

To establish property 2, suppose that y » yi(x) and y ** y 2 (x) are two 
solutions of (21-1). Then 

y% + pm + P2Vi 53 o, 

yl + pm + ?2V 2 s o. 

We must show that 

( 2 /i + y 2 ) n + Pi(yi + y 2 ) f + P 2 (yi + 1 / 2 ) 35 0. 


( 21 - 5 ) 



ORDINARY DIFFERENTIAL EQUATIONS 


52 


[chap, 1 


Inasmuch as the derivative of the sum of two functions is the sum of the de- 
rivatives, we can rewrite the left-hand member in (21-5) as 

(Vi + PiVi + Mi) + (vl + PiVz + V2V2)i 


and this vanishes by (21-4). 

Let us suppose now that by some means we have obtained two solutions 
yi(x) and y 2 {x) of (21-1). Then by the foregoing 

y(x) ~ cmix) + c 2 y 2 (x) (21-6) 


is a solution for every choice of the constants c x . We say that (21-6) is 
a general solution of (21-1) provided that for a suitable choice of the con- 
stants C{ the solution satisfies arbitrarily specified initial conditions, 

2/(*o) = 2/o, */'(«** o) * iL (21-7) 

To determine the restrictions on y\{x) and y 2 (x) ensuring that the solution 
(21-6) is, indeed, general, we insert (21-6) in (21-7) and obtain two linear 
algebraic equations, 

cm(x 0 ) + e 2 y 2 (x o) * y 0 , 

(21-8) 

cmM + c 2 y 2 (x () ) * y (h 


for Ci and c 2 . The system (21-8) can be solved for Cj and c 2 (for arbitrarily 
specified Xo, y 0) and y () ) if, and only if, the determinant 


W{y u y 2 ) s 


Vx(x) y 2 {x) 

y[(x) y 2 (x) 


9* 0 


(21-9) 


for every x = x 0 in the interval. If W(y lt y 2 ) = 0 for some value of x, 
the constants c t cannot be determined for every choice of y 0 and y r 0 and the 
solution (21-6) is not general. The determinant W(y X} y 2 ) is called the 
Wronskian after the Polish mathematician G. Wronski, who deduced 
the criterion (21-9) for the generality of solution (21-6). 

The condition (21-9) is equivalent to the statement that the solutions 
yi(x) and y 2 {x) are linearly independent. We say that yi(x) and y 2 (x ) 
are linearly independent if the identity 

cm(x) + c 2 y 2 (x) m 0 (21-10) 

can be satisfied only by choosing ci » c 2 = 0. When nonzero constants 
Ci and c% can be found such that c x yi(x) + c 2 y 2 (x) e? 0, we say that y x (x) 
and y 2 (x) are linearly dependent. In other words, linear independence of 
yi(x) and y 2 (x) means that the ratio y 2 (x)/yi(x) is not a constant. But if 
this ratio is not a constant, its derivative 

Vm - V\V% 

V\ 


(21-11) 



SEC* 21] LINEAR DIFFERENTIAL EQUATIONS 53 

is not identically zero. We note that the numerator in ( 21 - 11 ) is precisely 
W(i /t,y 2 ) ** VMi ~~ ViV 2* We have shown that if the solutions yi and y 2 
are linearly independent, then WO/ 1 , 1 / 2 ) 5 ** 0 for some value of x. Con- 
versely, if W ( 2 / 1 , y 2 ) 1=5 0 for some x « x 0 , so that 

Vi(. x o) Vn(* 0 ) = 

V\ (*o) v'i(x 0 ) 


we can show that the solutions y x (x) and y 2 (x) are linearly dependent , for 
we can choose nonzero constants c x and c 2 in (21-6) so that at the given 
point x «= z 0 our solution satislies the initial conditions 

y(*o) ’= 0, y'(x 0) - 0. (21-12) 

But if ?/(x) in (21-6) satisfies these conditions, then i/(x) ss 0 because there 
exists only one solution of Eq. (21-1) satisfying initial conditions (21-12) 
and a solution ?y(x) = 0 obviously satisfies these conditions. We have 
thus shown that the nonzero constants c* and c 2 can be found such that 
C\y\(x) + c 2 y> 2 (x) ~ 0 for all values of x, and hence the solutions y x (x) 
and ?/ 2 (x) are linearly dependent. 1 

It follows from this that the problem of finding the general solution of 
(21-1) reduces to the search for some pair of linearly independent solutions 
y i(x), i/ 2 (x). ft should be remarked that no formula is available for the 
determination of solutions of the general second-order linear equation 
In the special instance when the coefficients p x and p 2 in (21-1) are con- 
stants, the general solution, as we shall see in the following section, is 
deduced easily. 

Example: Verify that 

y * q sin x + c 2 cos x 

is the general solution of 

v" + y « 0 

and determine the particular solution such that 

1/(0) - 1, y\ 0) - M. (21-15) 

The fact that y x = sin x and y 2 «* cos x are, indeed, solutions of (21-14) is easily verified 
by substituting y » sin a: and y == cos x m (21-14). Hence their linear combination 
(21-13) is a general solution provided that the determinant (21-9) does not vanish. In 
our case, 

W(y h y 2 ) - 

and thus (21-13) is the general solution. To determine the constants a such that the 
solution satisfies conditions (21-15), we form the set of Eqs. (21-8), 

ci sin 0 4" q cos 0*1, 

cj cos 0 - c 2 sin 0 » 


sm x cos x 
cos x — sin x 


(21-13) 

(21-14) 


1 See in this connection Prob. 0, Sec. 21. 



54 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

from which it follows that c\ *■ es ** L Thus the desired solution is y m sin % 4 

COS X. 

PROBLEMS 


i* Verify that y ** e* and y — e~ z are linearly independent solutions of y" — y « 0, 
Also show that y\ ■» sinh x and y% * cosh x are a pair of linearly independent solutions 
of this equation. 

2. Show that y * cie* 4- is the general solution of y" — 3 y f 4 2y » 0, and find 
the solution satisfying the conditions t/(0) ** 0, y'(0) *» 1. What is the solution satisfying 

m - V'( 0) - 0? 

8. Show that yi « l/x and 3/2 ** x 6 are linearly independent solutions of x 2 y" — 
3 xy’ — by ** 0 if x ^ 0. 

4. Verify that y * cix 2 4 cz* is the general solution of x 2 y" — 2 xy f 4 2y ** 0 if 
x & 0, and find the solution such that y( 1) * 2 and y'(l) » 0. 

8. Show that y ■* c.\e} x 4 cjx/? 2 * is the general solution of y" — 4</' 4 4y ** 0, and 
find the solution for which y(0) «* 1, y'( 0) = 4. Also find the solution such that y(0) ** 
V'( 0) * 0. 

0. Compute the derivative of Wiyuy*) * 1/13/2 ~ VzVi, where yi(x) and y 2 ,(x) are two 
solutions of (21-1). Show that jqj/2 — yiyi 4 Pi(x)(yiV 2 — VtV\) ** 0 and dW /dx 4 

Pi (x)W «* 0. Thus W(yx,y 2 ) ** WoC~^ x o Pl( * >d *, where Wo is the value of W(yi,yt) at 
x « xo. Conclude from this that if W(y\,y 2 ) does not vanish at x «■ xo, it does not 
vanish for any value of x. This result is known as Abel's theorem. 

22 . Homogeneous Second-order Linear Equations with Constant Coef- 
ficients. Consider the equation 

y" + Pi V 9 + V 2 V * 0 (22-1) 

with constant coefficients pi, p 2 . If we substitute 

3/ - (22-2) 

in (22-1) and note that y r = rwe mx , y" = m 2 e mx , we obtain the equation 
(m 2 4 Pim 4 p 2 )e mx » 0, 


or 


m 2 4 Pim + p 2 = 0, 


(22-3) 


since e mx 9 * 0. Thus, if m in (22-2) is chosen as a root of the characteristic 
equation (22-3), then (22-2) will be a solution of the given equation. The 
roots of the quadratic equation (22-3) are 


-Pi db Vp? - 4 p 2 

m 

2 

If pi — 4p 2 > 0, there will be two distinct real roots, m 
«* m 2 . In this event, 

y = y ** 


(22-4) 


ini and m 


are a pair of linearly independent solutions of Eq. (22-1), since 



SEC. 22} LINEAR DIFFERENTIAL EQUATIONS 55 

is not a Constant when m x ^ m 2 . Hence if m x ^ ra 2 , the general solution 
of (224) is 

y * Cl e M i x + c 2 e m **. (22-5) 

If pf — 4p 2 < 0, the roots (22-4) are conjugate complex numbers , 
mi »fl + bi, m 2 = a — 6i, 


and the complex functions 

2/i = e (a+6,)I , 2/2 = (22-6) 

are linearly independent solutions of (22-1), We can write (22-6) in a 
trigonometric form with the aid of Euler’s formula [cf. Eq. (17-3), Chap. 2] 

e (a±bi)x 0<w( COR ( )X -k { s J n 


so that ?/i = e ax (cos 6x + i sin fcr), 

7/2 — e aar (cos br — z sin 6x), 


(22-7) 


are the complex solutions of (22-1). 

We show next that when Eq. (22-1) with real coefficients p } and p 2 has 
a complex solution of the form y = n + zV, then the real functions u and 
v are solutions of this equation. Indeed, the substitution of y — u + iv in 
(22-1) yields on rearrangement 


(u" + p x u f + p 2 u) + z( v" + p x v' + P 2 V) * 0, 


and this can vanish if, and only if, 

u" + p x u' + p 2 u - 0, 


v" + piv' + p 2 v = 0. 

Thus y — u and y =* v satisfy (22-1). 

Referring to (22-7) we see that corresponding to a pair of complex roots 
m = a ± bi of th'» characteristic equation, we have a pair of linearly in- 
dependent real solutions 

7/i = e ax cos bx, t/ 2 ” c™ sin bx. (22-8) 

It remains to consider the case when p\ — 4p 2 = 0. In this event the 
characteristic equation (22-3) has a double root 

~~Pi 

m x = m 2 — — — * 


and the foregoing method yields just one distinct solution y x *= e mx f with 
m ** —pi/2. We can verify by direct substitution that another solution 
is 1/2 xe mx , which is obviously linearly independent, since y 2 /y x 335 



ORDINARY DIFFERENTIAL EQUATIONS 


v 


56 


(chap* 1 


x 9 * const. Thus, when the characteristic equation has a double root 
m a* — px/2, the general solution of (22-1) is 


y = (cj + c 2 x)e m * (22-9) 

In the following section we deduce this solution with the aid of the useful 
notion of differential operators, which will be of help in resolving the cor- 
responding situation involving multiple roots of linear equations with 
constant coefficients of order higher than 2. 

We illustrate the foregoing discussion by examples. 


Example 1. Find two linearly independent solutions of y" 4- 3 y' -f 2y ** 0, and thus 
obtain the general solution. Referring to (22-3), we see that the characteristic equation 
in this case is 

m 2 4- 3m + 2 - 0, (22-10) 


which, on factoring, yields 

(to + 1)(to + 2) - 0. 


Thus, the roots of (22-10) are wi ■* —l, *■ — 2, and hence the general solution is 

y » Cie“" x 4* C 2 < ? ‘“ 2;c . 

Example 2. Solve y" + -4- 5 v «* 0. The characteristic equation is 

m 2 + 2m + 5 « 0, 

and hence 


m 


-2 dt VT - 20 


-1 dr 2 i. 


Accordingly, the complex solutions are 

Vx » «<-*+*>* t/ 2 - e ( ~ l ~ 2l)ar , ( 22 - 11 ) 

and by (22-8) the linearly independent real solutions are 

yi = cos 2.r, y% ~ e~ z sin 2x. (22-12) 


It should be remarked that for many purposes the complex form of solutions (22-11) 
is just as useful as the real form (22-12). 

Example 3. Solve y ,f 4* 2/ 4- y ** 0. The characteristic equation 

m ? 4- 2m -f 1 *=0 


has a double root m * — 1. Accordingly, a pair of linearly independent solutions of the 
given equation is y\ » e~~ x , yi ** j-e“ x . By (22-9) the general solution is 

K“(q4 C2x)e~~ x . 


PROBLEMS 


Find the general solutions of; 

i. y” 4- 3j/' - 54 y » 0; 

3 . y" - 2y' 4- y « 0; 

A | /" + 4y - 0; 

7. y" - 4* hy - 0. 


2. y" - by' + 6y * 0; 
4. y" — 4y « 0; 

6. y" - 4y' + 4y « 0; 



SBC, 23] UNEAR DIFFERENTIAL EQUATIONS 57 

23, Differential Operators. We introduce a new notation for the deriv- 
ative symbol and write D as d/dx and, more generally, D n m d n /dx 1t . 
Thus, D sin x means ( d sin x)/dx ~ cos x, and D 2 sin x » (d 2 sin x)/dx 2 
= d/dx (d sin x)/dx * —sin x . Since 


dcu(x) du 

« C — 

dx dx 

and 

d(u + t>) 

dx 

du dv 

dx dx 


we see that 





Dcu(x) *s cDu 

and 

IJ(u + v) 

= Du + Dv. 

(23-1) 

Moreover, since 

cT dy 

d nJrl y 




dx n dx 

dx n+1 ' 



we can write 

D n (Dy) = 

■■ n n+ 'y. 


(23-2) 


We agree also that 


(D + m)y e [)y + my. 


If the symbol (D + Wj)(D + m 2 )?/, where m x and m 2 are constants, is 
interpreted to mean that (D + ?^i) operates on (D + m 2 )y s (dy/dx) 
+ m 2 y, we find that 

(Z> + wj)fD + m 2 )y » [Z> 2 + (wi + m 2 )0 + mim 2 ]y. (23-3) 
From the structure of the right-hand member of (23-3) it follows that 

(D + m])(I) + m 2 )y — (I) + m 2 )(D + mi)y. (23-4) 

Making use of these properties, we can write Eq. (22-1), namely, 
d 2 y dy 

~~ 2 ' + Pi — + P2V — 0, (23-5) 

as (Z> 2 + PiD + p 2 )y « 0, (23-6) 

in which the differential operator 

D 2 + piD + p 2 

behaves as though it were an algebraic polynomial. 

We observe that this polynomial is identical with the polynomial in the 
characteristic equation (22-3). On noting (23-3), we see that (23-6) can 
be written in factored form as 


(D — m x ){D — m 2 )y « 0, (23-7) 

where rn x and m 2 are the roots of (22-3). Now, if m x ^ m 2 , the general 
solution of (23-7), as shown in the preceding section, is 

y mm + C 2 e m * X . 



ORDINARY DIFFERENTIAL EQUATIONS 


58 

To obtain the general solution of (23-7) when mi 
as follows: We set in (23-7) 

(D - m)y « v 

and obtain a first-order equation for v , 

(D — m)v — 0, 
dv 

or mv — 0. 

dx 


[chaf. 1 
m 2 ® m, we proceed 

(23-8) 

(23-9) 


Its general solution is v = cie mjr . The substitution of this in the right- 
hand member of (23-8) yields the first-order linear equation for y 

dy 

my « Cie tnx t 

dx 

whose general solution is easily found from (10-8). We thus get the solution 

y * c x e mx + c 2 xe mT y (23-10) 

which agrees with (22-9). 

Example 1. Find the general solution of y" -f 5 y' -f 6y ® 0, This equation can be 
written as 

(D 2 + 5D + 6)2/ - 0. 

On factoring the operator we get 

(D + 2 )(D + 3)2/ - 0, 

and thus the general solution is 

y =» cie"" 2 * 4- C 2 # , " 3x . 

Example 2. Solve 2 /” — 4y' -f « 0. We write this equation as 

(Z) 2 - 4Z> + 4)?/ * 0 
or (D - 2){D - 2)?/ - 0. 

Since the roots of the characteristic equation are equal, the general solution is 

y « ae u + cjxe 2 *. 


PROBLEMS 

Solve: 

1. - HD ~ M)V - 0; 

3. (Z>* + D - 2)2/ - 0; 

5. (D*42 D + 1)|/ - 0. 


2. (D* - l)y - 0; 
4 . (D - 3) J i/ - 0; 



sec, 24] 


LINEAR DIFFERENTIAL EQUATIONS 


59 


24. Nonhomogeneous Second-order Linear Equations. The equation 

y" + Pi(z)y' + p 2 &)y - /(«), (24-i) 

in which the right-hand member f(x) is a known continuous function, is 
called nonhomogeneous. The existence and uniqueness theorem of Sec. 1 
guarantees that this equation has one, and only one, solution satisfying 
the conditions 

y( x o) = yo, y'(x 0 ) = vi 

whenever the coefficients Pi(x) are continuous functions. If y » u(x) 
is any solution of (24-1) and y x (x) and y 2 (x) are linearly independent 
solutions of the associated homogeneous equation 

y" + Pi(*)y' + P 2 (x)y = 0, (24-2) 

the general solution of (24-1) is 

V « c x yi(x) + c 2 y 2 (x ) + u(x). (24-3) 

The fact that (24-3) is, indeed, a solution of (24-1) follows upon substitut- 
ing (24-3) in (24-1) and noting that 

u"(x) + p x (x)u'(x) + p 2 (x)u(x) = f(x) 

and that y = Ciy x (x) + c 2 y 2 (x) satisfies the homogeneous equation (24-2). 
The proof that (24-3) is the general solution is virtually identical with the 
proof in Sec. 21 for the homogeneous equation. 1 

We shall see in Sec. 28 that a particular integral u(x) of (24-1) can al- 
ways be determined whenever the general solution 2 of the associated homo- 
geneous equation (24-2) is known . In special instances, however, particular 
integrals of nonhomogeneous equations can be deduced without the 
knowledge of the general solution of the homogeneous equation. This 
vsimpler technique, based on judicious guesses of the probable forms of 
particular integrals, is known as the method of undetermined coefficients. 
It is applicable to linear equations with constant coefficients only when the 
right-hand member f(x ) has certain special simple forms. 

We illustrate the essence of this method by several examples and develop 
it in greater detail in the following section. 

Example 1. The right-hand member of 

y" + 3 y' + 2y-2e* (24-4) 

suggests that it probably has a solution of the form y * ae x , for the differentiation of 

1 The only difference is that the termB yo — and yj — uo(xo) instead of yo and yo 
now appear in the right-hand members of Eqs. (21-8) 

* This general solution is often called the complementary function. 



60 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

exponentials yields exponentials. Accordingly, we take y ~ ae* as our trial solution, 
substitute it in (24-4), and obtain 



os* 4- 3oe* 4- 2oe* - 2c*. 


On dividing by e x we get 

6a » 2 or a }4- 


Thus, y ** \ie s is a solution of (24-4). The characteristic equation of the associated 
homogeneous equation 

V" + 3y' 4- 2y - 0 (24-5) 

is 

nr ^ 3m -f 2 *= 0 


or 

(m 4* l)(m 4- 2) =» 0. 



Hence ft pair of linearly independent solutions of (24-5) is y * e~ x , y — e~ 2x , and the 
general solution of the given equation (24-4) is y =* C\e~ x -f cae'"' 2 * 4* 

If the solution of (24-4) satisfying the given initial conditions is required, the con- 
stants a must be determined from these conditions. For example, if we seek the solu- 
tion such that y( 0) » and y'( 0) =* 0, we obtain, on setting x « 0 in the general 
solution, 

— % *■ Cl + Cj -f H, 

or ci 4- Oi — — 1. 

Also, y' «■ — cic“* — 2 c 2 e~ 2 * + 

and since y'(0) «* 0, we have 

0 * —ci — 2 c 2 -f 

cr ci + 2 cj * 

We easily verify that ci — — c 2 = 4-46, and the desired solution therefore is 

V - + He*. 

Example 2. If we attempt to obtain a solution of 

v" + 3y' 4- 2y « 2c"** (24-6) 

by taking a trial solution y — ac~ x we get 

ac“ x — 3ac~ x 4* 2ae~ x 2e“ x . 

This gives a nonsensical result, 0 = 2e“ x . The reason that the trial solution of the form 
y m ae~ x is not suitable in this case is the following: The homogeneous equation asso- 
ciated with (24-6), as we saw in the preceding example, has a solution y * ac~ x , and the 
substitution of it in (24-6) naturally makes its left-hand member vanish. In this case 
we take the trial solution in the form y =* axe~ x . Then, y' «* m ~~ 9 — axe~ x } y” ■> 
— as~* — ae~ x 4* axc“ x , and the substitution in (24-6) now yields 

— 2a«“ x 4* axe~~ x 4- Sae^ - 3axe“* 4- 2 axe~* - 2e~*, 

or ac*~ x *» 2«~ x . 

Thus a ** 2, and a solution of (24-6) is y - 2ze“ x . The general solution of (24-6), there- 
fore, is 

y m ci«“ x 4- c*r u 4- 2xe~ x . 



BBC. 24] UNBAR DIFFERENTIAL EQUATIONS 61 

Example 3. Find the general solution of 

y" 4 V + V « (24-7) 

We recall (Example 3, Sec. 22) that a pair of linearly independent solutions of the asso- 
ciated homogeneous equation is y » e~*, y *» zf~ x . Accordingly, neither y *■> oe~* nor 
y » axe" 1 is suitable as a trial solution of (24-7). In this case we take the trial solution 

V « ax 2 e~ x . (24-3) 

We compute 

y’ = 2axe~ x — a£ 2 e~ x , 

2/" «* 2a^~ s — 2axe~ x — 4- ax 2 e~* 


and on making substitutions in (24-7) find 

2oe~ x — 4axe~ T 4 ax 2 e~ x 4 4aj:s“ x — 2 ax 2 e~ z 4 ax 2 e~ x ** e ~* 


or 2a« x « e~ z . 

Thus, a » J^S, and from (24-8) 3/ *» ^x 2 c“ z is a solution of (24-7). Its general solution 
is 

y = cie~~ x 4 tyxer* 4 Vzx 2 e'~ r . 

These examples suggest a procedure to be followed in obtaining particular 
integrals of equations with constant coefficients of the type 

y" + Pi?/ + P 2 ?y = Ae kx . (24-9) 

The characteristic equation associated with (24-9) is 

to 2 + pi m + p 2 = 0. (24-10) 

If this equation has two distinct roots m = rrti and m = m 2 , then the 
linearly independent solutions of the homogeneous equation 

V " + Pi2/' + V 2 V = o (24-11) 

are y = and ?/ = e” 1 * 7- . When Eq. (24-10) has a double root m 2 = m 1? 
the linearly independent solutions of (24-11) are y = e w i x and ?/ = 

Now, if A: in the right-hand member of (24-9) is not equal to either mj or 
m 2 , Eq. (24-9) has a solution of the form y = ae kx . If A: is a simple root 
of (24-10), then (24-9) has a solution of the form y = axe kx . When A; is a 
double root of (24-10), the particular integral can be taken in the form 
y * ax 2 e kx . 

Similar considerations apply to equations of the form 

y" + Ply' + P 2 V = Ao + A X X 4 h A n x n . (24-12) 

If 1 pa 5 ^ 0, we can take the trial solution 

y « a 0 + &\x + * * • + ®nX n • (24-13) 


1 This means that m » 0 is not a root of the characteristic equation (24-10) and hence 
(24-11) has no solution y — const. 



62 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

The substitution of (24-13) in (24-12) then yields on comparison of like 
powers of x on both sides of the resulting equation the values of the un- 
known constants a*. If P 2 * 0, the characteristic equation (24-10) has 
m m 0 as one of its roots. In this event the trial solution can be taken in 
the form 

y ** x(a 0 + aix H f* a n x n ). (24-14) 

We illustrate the use of these rules by two examples. 

Example 4. Find a solution of 

y" + Sy f + 2y - 1 + 2x. (24-15) 

Since p% -* 2 5* 0, we take the trial solution 

y « oo + a } x, (24-10) 

substitute it in (24-15), and find 

3ai -f 2(ao -f «ix) 1 -f 2x. 

On comparing like powers of x, we get 

3a 1 *4 2ao ** 1, 2ai ** 2, 

whence 

a* ■» 1, oo *» — 1. 

The substitution of these values in (24-16) gives the desired solution y « —I 4- x. 
Example 5. Find the solution of 

y" + 3 / = 1 - 9x 2 (24-17) 

satisfying the conditions y( 0) « 0, y'( 0) * 1 . 

Since p% ** 0 in (24-17), we seek a solution in the form 

y * x(ao + ojx 4~ a 2 x 2 ), (24-18) 

We compute 

y' ** oo 4* 2ojx 4- 3 o2X 2 , 
y" ** 2oi 4* 

and insert in (24-17). The result is 

2ai 4- 6 oax 4* 3(ao 4- 2aix 4- 3aax 2 ) ■» 1 - 9x 2 
or 2«i 4- 3ao 4- (602 4- 6«i)x 4- 9o*x 2 « 1 — 9x*. 

Hence 

2oj 4- 3oo ® 1, 

O02 4" ** 0, 

9os *• —9. 

Solving these equations, we get 

«2 -1, * 1, oo » 

and the substitution of these values in (24-18) gives 

y - x(-% -f x - x 8 ). 



SEC. 24] LINEAR DIFFERENTIAL EQUATIONS 63 

The characteristic equation for (24-17) is 

m 2 -f 3m =* 0. 

Since its roots are m * 0, m » —3, the general solution of (24-17) is 

V “ O + c**-’ 1 + *( - H + x - I s ). (24-19) 

To determine the constants so that the solution (24-19) satisfies the given conditions, 
we compute 

y'(x) - — 3c 2 e- s * - (% - 2x + 3i s ). 

The conditions y( 0) * 0 and y'( 0) ■» 1 then demand that 

n + fj « 0, 

~3c 2 - H • 1. 

Thus, C2 » — c\ * and hence the desired solution is 

V m % ~ %€ ~ u -f x(~M + I - X 3 ). 

We state in conclusion that the trial solution for the more general 
equation 

y" + Pi y' + P 2 P = e kx (A 0 + A IX H 1- A n x T ‘) (24-20) 

can be sought in the form 

V = e kT (ao 4- a x x 4 b a n x n ) (24-21) 

if k is not a root of (24-10). If k is a simple root of (24-10), the trial solution 
(24-21) must be multiplied by x and, if the root is double, by x 2 . 

PROBLEMS 


Obtain the general solution: 


1. 

y" - 

W + 63/ = 

e 4r ; 

2. y" 

4- 2 y' 4 y « x; 


3 . 

v" + 

57/' 4- 6j/ * 

«*; 

4. y" 

— 2i/' 4- y «* x; 


6. 

(O 2 - 

- 1 )y « 5x - 

- 2 ; 

6 . (O 2 

- 1)?/ - e 2 *(x - 

- i); 

7 . 

(O - 

1) 2 ?/ - 


8 . (O 2 

~ 6/> 4- 9)3/ « 

c 3 *; 

9. 

0(0 

+ «)lf - 3; 


10. y” 

+ 92 / - x 2 - 2 x 

4- 1; 

11 . 

f"- 

y ** 


12. y" 

4 - ?/ *= x 3 -b x; 


13 . 

ll" - 

W 4* 6 y - 

ae** 1 *; 

14. (O 

-- 1) 2 t/ * c*(x - 

i); 

15 . 

(O 2 - 

- 5D + 6 ) 3 / 

= 3x s + 4i - 2; 

16 . (D 2 

- 5D)y * 3x 2 • 

f 4* 


Obtain the solution for each of the following equations satisfying the given conditions: 

17 . y" + by' + iy - 20c 1 , ?/(0) = 0, y'(0) 2; 

18 . y" + y' = 1 + 2x, 3/(0) - 0, t/'(0) = 0; 

18 . y" + y’ -0, y(0) - 0, j/'(0) - 0; 

20. y" + 4j/' + 3j/ - x, y(0) = — 46, J/'(0) - %\ 

21. y" + 4 y' + 3y - 0, y(0) - 0, j/'(0) - 0; 

22. j/" + 4y' + 3y - y(0) - 1, y'(0) - 0. 



ORDINARY DIFFERENTIAL EQUATIONS 


[CRAF. 1 

25. The Use of Complex Forms of Solutions in Evaluating Particular 
Integrals. The method of determining particular integrals of Eq. (24-9), 
described in the preceding section, can be extended to equations of the 
form 

y" + piy' + P 2 V « Ae kx cos nx f (25-1) 

V n + PiV f + PiV = Ae kx sin nx } (25-2) 

in which k may be equal to zero. If we recall the formula 

e %nx a cos nx + i sin nx, 

it becomes clear that Ae kx cos nx and Ae kx sin nx are, respectively, the real 
and imaginary parts of the function Ae ik+in)x . Now, if instead of Eqs. 
(25-1) and (25-2) we consider the equation 

y" + Piy ' + P2V « Ae {k+%n)z (25-3) 

and obtain its solution y * u + iv, the real part u of such a solution will 
satisfy Eq. (25-1) and the imaginary part v will be a solution of (25-2). 
We illustrate this method of deducing solutions of equations in the forms 
(25-1) and (25-2) by examples. 

Example 1. Find a solution of 

y" + y -» 3 sin 2x. (25-4) 

Since e i2x ■» cos 2x + i sin 2x, we consider, instead of (25-4), the equation 

y” + V ** 3(cos 2x -f- i sin 2x) =* 3e 2tx . (25-5) 

The imaginary part of a solution of (25-5) is clearly a solution of (25-4). Equation (25-5) 
has the form (24-9) with k « 2t, and since neither of the roots of the characteristic equa- 
tion m 2 *f 1 *■ 0 is equal to 2i, we take the trial solution 

y - ae 2l *. 

Now y' « 2 iae 2tx , 

y” = (2 i) 2 ae 2lx «* ~iae 2vx , 
and the substitution in (25-5) yields 

~4ae 2ix -f ae 2tx ** 3e 2t *. 

Thus, a « — 1, and consequently y = -c 2u: is an integral of (25-5). The imaginary 
part of — e 2tx is — sin 2x, and hence a solution of (25-4) is y « — sin 2x. 

Example 2. Find one integral of 

y” *f y * 3 cos x. (25-6) 

Since e** ■* cos x -f t sin x, we consider 

y" + y — 3(cos x -f i sin x) m Ze w f (25-7) 

the real part of the solution of which satisfies (25-6). This time k in (24-9) is 4-t, and 



65 


SBC. 25 ] LINEAR DIFFERENTIAL EQUATIONS 

since the roots of the characteristic equation are we take the trial solution 

y « axe**. 


From (25-8), 

y' « ae** -f aixe**, 
y" « 2aie^ - axe <*, 


and the substitution in (25-7) gives 

2aie tx ~ axe** ~f- axe** «» 3e**. 


(25-8) 


Thus, a *« 3/2i ** — %i, and therefore 

y m — %ixe** m — %ix ( cos i + i sin x) 

is a solution of (25-7). The real part of this solution is %x sin x, and we conclude that 
y ** Y$x sin x is a solution of (25-6). 

Example 3. Fmd a solution of 


y" 4 2 y' 4 2y * e~~ x cos x. (25-9) 

Since e~ * cos x is the real part of 

e x (cos x 4 i sin x) ** e~~ r e tT « e x( ~ 1+ ‘\ 
we consider the equation 

y" 4 2 y f 4 2 y * e* ( “ 1+,) . (25-10) 

The roots of the characteristic equation 

w 2 4 2m 4 2*0 

are m ® — 1 db t, and since one of these roots appears in the exponent in (25-10), we 
take the trial solution 


- ox^~ 1+< >. 


Then, 


y' =» ae x( " 1+t > 4 ox( — 1 4 t)c x( ~ 1+t) , 
y" « 2a(-l 4 4 ax( — 1 4 i)*e x( ~ 1+i) t 

and on making substitutions in (25-10), we find 

2 cue x ^~ 1 ‘*' t ^ «= 

1 


so that 

Thus an integral of (25-10) is 
1 


2 i 


y «*» — xe x( 1+v) ** — : xe” x (cos x 4 t sin x). 

2i 2t 

The real part of this, >£xe~ x sin x, is a solution of (25-9). 

The methods of this and the preceding section can be extended to 
equations 

v" + piy' + ? 2 V = /(*) ( 25 - 11 ) 

in which the right-hand member is a sum of several functions of the types 



66 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

considered in these sections, for suppose that f(x) « f x (x) + / 2 (x), so that 
(25-11) reads 

V " + PiV' + PnV - /i(*) + / 2 (x). (25-12) 

If we consider a pair of equations 

y" + Piy' + p 2 y = Mx), 

(25-13) 

2/" + PiV 1 + P2V = / 2 (x), 

and denote the solution of the first of these by y = tq(x) and that of the 
second by y = %(x), then y = n x (:r) + n 2 (x) will be a solution of (25-12). 
The proof follows at once on inserting y — n^x) + u> z (x) in (25-12). As 
an illustration of the use of t his theorem we consider a simple example. 

Example 4. Find one solution of 

y" ~b y “ 3 cos x + 1 + 2e*. 

We consider three equations: 

y " -f y » 3 cos x, 
y" + y - 1, 
y" + y - 2e*. 

A particular integral of the first of these, as shown in Example 2, is y ® %x sin x, and 
solutions of the second and third equations are, respectively, y ** 1 and y » e*, as is 
clear by inspection. Hence an integral of the given equation is y » %x sin i + 1 + c r . 


PROBLEMS 

Solve: 

1. (Z> 2 - 32> + 2)y - cos 2x; 

2. (Z) 2 *4* 4)y * cos 3x; 

3. ( D 2 - 3 - H)!/ * — cos x — 3 sin x; 

4. y" + 5y' -f 6y * 3e“ 2x + *' 3x ; 

5. y " + 2y' 5y » e 1 sin 2x; 

8. y" — y' — 6y = e 3x cos 3x; 

7. (Z> 2 - 25)y - e 5x + x 2 - 4x; 

8 . (D 2 + l)y 3 sin 2x - 9 cos 3x. 

Obtain the solution satisfying the conditions y( 0) » 0, y'( 0) » 0 for each of the fol- 
lowing: 

9. y" ~ y - sin x; 10 . y" -f 2y' -f 5y = 0; 

11 . y" — 2y' ® e“ x cos x; 12 . y" -f- y ® cos x + 1. 

26. Linear nth-order Equations with Constant Coefficients. The results 
of Secs. 22 to 25 are easily extended to nth-order linear equations 

y in) + P\y {n ~ l) H f p„y = /(*) (26-1) 

with constant coefficients. In dealing with such equations it is convenient 



SEC. 26] LINEAR DIFFERENTIAL EQUATIONS 67 

to make a systematic use of the operator notation introduced in Sec. 23 
and write (26-1) in the form 

(D n + Pi D n *~ l H b p n -\D + p n )y = f(x). (26-2) 

The homogeneous equation associated with (26-2) is 

(D n + V\D n ~ l + • • • + pn^D + p n )y » 0, (26-3) 

and if one substitutes in it y = e mr , there results 

(m n + pirn*-" 1 H b Pn-iw + Pn)c*” r = 0. 

It follows that ?/ ~ e mx is a solution of (26-3) whenever m is a root of the 
characteristic equation 

m n + pim n ~ l H b Pn-\m + p n = 0. (26-4) 

If this equation has n distinct roots 

m ~ m u rn — m 2 , . . m ~ m n , 

then 2 / = e w > x , 2 / = c w ’ x , . . . , y = e m - x 

are distinct solutions, and we can conclude (see Sec. 27) that 

y = c,c m > x + c 2 c w 3 x H b c n c w « x (26-5) 

is a general solution in the sense that the arbitrary constants c t in (26-5) 
can be determined to satisfy the prescribed initial conditions 

y(*o) = yo> y'(x o) = 2/o, • • i/ ln ” l) (zo) = (26-6) 

Since the coefficients m (26-4) are real, the complex roots of (26-4) must 
necessarily occur in conjugate pairs. Thus, if nil ~ a + to and m 2 — 
a — to are a pair of such roots, the solutions corresponding to them are 

y x = *>(*+&»)* a* e ax (cos + t sin 6x), 

t/ 2 = f (a-bi)x _ e ax^ cos fa _ ^ g j n 

As in Sec. 22, we prove that the real and imaginary parts of these solutions 
yield a pair of linearly independent real solutions 

y = e ax cos bx ) y * e ax sin bx. 

When the roots of characteristic equation (26-4) are not simple, and if, 
for example, the root mi has the multiplicity k , then corresponding to it 



68 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

there will be a set of k distinct solutions, 1 

! /i -*"»•, y 2 = xe^ x , y k = x k ~'e m i x . (26-7) 

The proof of this assertion follows upon making obvious modifications in 
the argument presented in Sec. 23. 

We illustrate these statements by two examples. 

Example 1. Find the general solution of the fourth-order equation 

3^ - 2 r 4- 2 y" - 2y' + y * 0, (26-8) 

or (D 4 - 2D 8 4- 2D 2 - 2D 4- 1 )y * 0. (26-9) 

The characteristic equation for (26-8) has the structure determined by the operator in 
(26-9). It is 

m 4 - 2m 8 4* 2 nr - 2m 41=0. 

On factoring this we get 

(m 2 4- l)(m - l) 2 - 0. 

Thus, there are two simple roots mi = t, m 2 = —i and the double root m3 ■* m 4 ** 1. 
Solutions corresponding to these roots are 

3/1 « e 4 *, 3/2 * c""" 3/3 - e x , 2/4 * 

and the general solution is 

y *= 4* 4 C 3 C X 4 

This can be written in real form as 

y «* Ci cos 3; -j- C2 sin r + (C3 4- (\r)e x . 

Example 2. The equation 

(D 4 4* 3D 3 4- 3D 2 4 D)y « 0 
or D(IJ 4- 1 ) z y - 0 

has the characteristic equation 

m(m 4- l) 3 - 0. 

Accordingly, the general solution is 

y •» cie 0x 4* C 2 &~ x 4* r a zc“ x 4* r i i 2 e'" x . 

An argument in every respect similar to that given in Sec. 24 yields the 
result that when y = u(x) is any solution of Eq. (26-1) and y = c x y x (x) 

+ C 2 y%(x) 4 h c n 2/ n (^) is the general solution of the homogeneous 

equation (26-3), then the general solution of (26~1) is 

V = c il/i (x) + C 2 y 2 (x) -| 1 - C„y„(x) + u(x). (26-10) 

1 If the complex root m\ ■* a 4- bi is of multiplicity then corresponding to this root 
and to its conjugate m2 « a — bi, there will be a set of 2k real solutions: 

e ax cos bx, xe° z cos bx, . . jr*” 1 #** cos for, 

sin bx, xe? x sin bx, . . . , sin bx. 



SEC. 26] UNEAR DIFFERENTIAL EQUATIONS 60 

The calculation of particular solutions u(x) by the method of undetermined 
coefficients for functions f(x) of the type considered in Secs. 24 and 25 
follows, with obvious minor modifications, the pattern of those sections. 
Without further ado we illustrate the procedure by examples. 1 

Example 3. Find a solution of 

3 /"' 4- y " 4 2 y' -x J 43^41. (26-11) 

The left-hand member of this equation contains no y (that is, pa * 0). On recalling the 
statement made for Eq. (24-12), we take the trial solution 

y « x(cio 4 d\x 4* oax 2 ), 

On computing the first three derivative^ we obtain 

y' » ao 4 2a it 4 Za^x 1 , 

V" “ 2aj -f 6a 2 x, 

- 6a* 

Substitution in (20-11) then yields 

(2ao 4 2ai -f (>a 2 ) 4 (0a 2 4 4a pjr 4 0a 2 x 2 « x 2 -f 3a: 4 1- 


Hence 0a 2 ~ 1, 

0a 2 4 4a i = 3, 

2ao 4 2ai 4 0a. 2 ** 1 

ami we conclude that 

a 2 ** a 1 =^ 2 , ao — -- 1 £. 

Accordingly, y « r( — 4 Jyx 4 VfcJ 2 ) is a solution of (20-11). 

Example 4. Obtain the general solution of 

(/J 3 - 3D 2 4 2 l))y * 4 4- 60e 5x . (20-12) 

The characteristic equation for (20-12) is 

m 3 — 3m 2 4 2m «= m(m — l)(m — 2) ** 0. 


Thus, the general solution is 

y - ci 4 c 2 <4 4 c 3 c 2x 4- a(x), 

where u(x) is some integral of (20-12). To obtain a(j) it is simpler to add the particular 
integrals of 

{D z - 3 D* 4 2 D)y • 4, (20-13) 

(JO 8 - 3L> 2 4 2D)y - 60c 5 *. (2(4-14) 

For the first of these we take a trial solution y » ax. We find on inserting it in (20-13) 
that a w 2, so that p =» 2x is an integral of (20-13). The substitution of p » ae hr in 
(26-11) yieltls, after simple calculation, a «* 1; hence p * e 5 * is a solution of (20-14). 
Accordingly, an integral of (20-12) is y « 2x 4 c Cjr , and the desired general solution is 

3/ =» ci 4 c 2 e z 4 c'3C 2x 4 2x 4 e 5iC . 


1 A general method is presented in Sec. 28. 



70 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

Example 5. Obtain one solution of (D 3 -f 2D *f 7 )y « —24e*cos2x. The right- 
hand side is the real part of —24f T e 2ix . To solve 

(D 3 + 2£> 4* 7)y - — 24e T e 2u? - -24f< l+fl) * (26-15) 

try y « Substitution gives 

(D* 4* 2D + 7)oe (1 + 2 * )z « ((1 -f 2i) 8 + 2(1 + 2?*) + 7]ae (1 + 2l) * 

~ (-2 + 2i)ar (l+20z * -24e <l+ * l) * 


where the last equality is stated because we want to obtain a solution of (26-15). 
follows that 


—24 

-2 -f 2t 



12 1 4 - i 

1 — i 1 4* i 


60 4 0, 


It 


and hence a solution of (26-15) is 


y «* 6(1 4“ i)e {l ~* 2l)r •* 6(^(1 4“ *)(eos 2x 4~ ? sin 2x) (26-16) 

Since — 24e z cos2x is the real part of ~24e (1+2t)x , a solution of the original problem is 
found by taking the real part of y in (26-16). Thus, 


y = 6e x cos 2 j — 6^ sin 2x. 


PROBLEMS 

Find the general solutions: 

1. (D - 5)(2 1) + 3 )l)y - 0; 


3. (/> 3 4- 3D 2 -f 3 D + 1 ) y - 0; 

6. (Z) 3 - 2D 2 4- D)y ~ 0; 

7. (i> 4 - k*)y - 0; 

9. (D 3 — I> 2 4- 4Z))j/ * 4x 4- e*; 

11. (O - 1)(Z> - 2) 2 ?/ - x 2 ; 

13 . Find the solution of y"' 4~ 2y 1 
y’( 0) * 0, iy”(Q) ~ -25/2. 


2. (D 2 4- l)(/> 2 4- 21) + 5)v - 0; 

4 . (D* 4- 8)?/ - 0; 

6. (D 4 4- 3 D z 4- 3D 2 + D)y - 0; 

8. (/) 3 - 3D 2 4- 4)?/ - 0; 

10. (/> 4 4- Dv — 2 cos x; 

12 . (D 4- 1)(D ~ l)(/> ~ 2)y - 1 - c J 

y' — 2y « 2c " Sx 4- 4x 2 which satisfies #(0) 


27. General Linear Differential Equations of nth Order. It is not dif- 
ficult to extend the considerations of Sec. 21 to a homogeneous nth-order 
linear equation 

y in) + pi(x)y {n ~ 1) + p 2 (x)y in ~ 2) + f p n -i (x)y' + Pn{x)y = 0 (27-1) 


with variable coefficients p t (j). Word-for-word repetition of the argument 
used to establish properties 1 and 2 of Sec. 21 leads to the conclusion that 

V * c t yi(x) 4- c 2 y 2 (x) H h c k y k (x ) (27-2) 

is a solution of (27-1), for an arbitral choice of the constants c *•, whenever 
2 /i(x), y 2 i x )> • • *> 2/ifcW is a set of solutions of (27-1). 

A set of k such solutions is said to be linearly independent if the relation 


CiV\{x) + c 2 y 2 (x) H f c k y k (x) as 0 (27-3) 

holds only when c\ c 2 = •• * - c k =* 0. When a set of constants c*, 



8KC. 27] LINEAR DIFFERENTIAL EQUATIONS 71 

not all of which are zero, can be found such that Eq. (27-3) is true, the 
solutions yi(x) are linearly dependent. 

It foIloAvs from the existence and uniqueness theorem of Sec. 1 that 
Eq. (27-1) has exactly n linearly independent solutions y\{x), . * y n (x), 
so that 

y = Vl(x) + C 22 / 2 M H h c n y n (x) (27-4) 

is a general solution of (27-1). This solution is general in the sense that 
the constants c t in (27-4) can always be found, so that there is a unique 
solution of (27-1) for the arbitrarily specified initial values 

y(*o) = 2/o, y'(x o) = 2 / 0 , •••. y (n ~"(xo) = */o n_1) . (27-5) 


An argument analogous to that used to establish the condition (21-9) for 
linear independence of two solutions leads to the result that the set of n 
solutions { y t (x)), i ~ 1, n, is linearly independent if, and only if, 
the Wronskian determinant 




2/i 

2/2 

• • 2/n 

t 


/ 

Vi 

2/2 

• . 2/n 


,/n-l) 

„ .(n 

Vi 

v-2 

• • 2/i 


(27-6) 


dot's not vanish for any x in the interval where solutions are sought. 

In contradistinction to the case of linear equations with cor >tant co- 
efficients, no formulas are available for solving general linear equations 
with variable coefficients of order 2 or higher. Certain special types of 
such equations, however, have been studied extensively, and as shown 
in Chap 2, Sec 12, their solutions may be obtained as power series. 

Just as in Sec. 24, we can show that if y = u(x) is any solution of the 
nonhomogeneous equation 


y (n) + Pi (•*•)//" ,J H h Vn- lW/ + Pn(-r)?/ = /(*), (27-7) 


(hen y = r^y ,(.r) + r 2 y 2 (x) -f c n y„(x) + »(x) (27-8) 


is the general solution of (27-7) whenever the are linearly independent 
solutions of the homogeneous equation (27-1) The determination of 
particular integrals of (27-7), as we shall see in the next section, is a straight- 
forward process provided that the general solution of the associated 
homogeneous equation is known. 


Example 1. Show that the set of functions t/i » x, yz ** x 2 , y$ ** x 3 is linearly inde- 
pendent if x 9 * 0. The Wronskian (27-0) for this set of functions is 


1 2x 3 or 

0 2 (u 


2x 3 . 


W(j i h y2,yz) * 



72 ORDINARY DIFFERENTIAL EQUATIONS [CHAP, 1 

Since W(gn,yi,w) does not vanish as long as x & 0, this set is linearly independent in 
any interval that does not include / = 0, 

Example 2. Test for linear independence yi » x 2 -f 2 x, » x 8 4* x, yz » 2x 8 — x 2 . 
We compute the Wronskian for this set of functions: 


W(y h yz,yz) 


x 2 4* 2x x 8 *f x 2x 3 — x 2 
2x 4* 2 3x 2 -h 1 6x 2 - 2x 
2 Ox 12x — 2 


0 . 


Since W(yi,y 2 ,y a ) «* 0, the given set is linearly dependent. This implies that a set of 
constants Ci, cq, cz ) not all zero, can be found such that 


cm 4- r 2 ya 4 cm * 0. 

Tliis, in turn, means that at least one of these functions can be expressed linearly in 
terms of the remaining ones. In fact, it is easy to check that t/ a ** 2y<i — j/j. 


PROBLEMS 

Test for linear dependence the following sets of functions: 

1. c~ x , l, e*, sinh x; St. 1, sin x, cos x; 

3. x 2 — 2x 4- 5, 3x — 1, sin x; 4. (x 4- l) 2 , (x — X) 2 , 3x; 

5. e* x , e bx , e cz , a ^ b t 6 c ^ a; 6. e lx , sin x, cos x; 

7. e x , xe*, x 2 e*. 

28. Variation of Parameters. We proceed to show that a particular 
integral y « u(x) of every nth-order linear equation (27-7) can be cal- 
culated by the so-called method of variation of parameters whenever the 
general solution of the related homogeneous equation (27-1) is known. 

To make the procedure clear, we first develop it for the second-order 
equation 

y" + Pi(*)y' + P 2 (x)y - /(*) (28-1) 

and then extend it to the general case of Eq. (27-7). Let us suppose that 

V = cxyi(x) 4- c 2 y 2 (x) (28-2) 

is the general solution of the homogeneous equation 

y" + V\ (x)y* + PzWy = 0. (28-3) 

We shall attempt to find an integral of (28-1) in the form 

V * + v 2 (r)y 2 (x) t (28-4) 

obtained from (28-2) by replacing the constants c» by some unknown 
functions v t (z). 

If we substitute (28-4) in (28-1), we shall obtain one equation which 
imposes a condition to be satisfied by two unknown functions v x (x) and 
v#(x). Since one such condition does not determine the unknown functions. 



LINEAR DIFFERENTIAL EQUATIONS 


73 


me. 28 ] 

we need another equation relating v x and v 2 . We shall impose this second 
condition in a way that would tend to simplify the calculation of v x and v 2 . 
If we differentiate (28-4), we get 

y' * (viVi + v 2 y 2 ) + (v[yi 4* v 2 y 2 ). (28-5) 

Now, the calculation of y" will he materially simplified if Vi and v 2 are 
chosen so that the expression in the second parentheses in (28-5) vanishes. 
Accordingly, we set 

v\yi + v 2 y 2 * 0 (28-6) 

and take y ' ~ v\V\ + v 2 y 2 . (28-7) 

Then y" » v x y[ + v 2 y 2 + v\y\ + v 2 y 2 . (28-8) 

The substitution from (28-4), (28-7), and (28-8) in the original equation 
(28-1) yields, on rearrangement, 

vi(y" + VxVi + P 2 I/ 1 ) + v 2 (y 2 + V\V 2 + 7>2!/2> + v\y\ + v 2 y 2 * f(x). 

(28-9) 

But since y x and y 2 are known to satisfy (28-3), the expressions in the 
parentheses in (28-9) vanish. We thus get 

vWi + v 2 y 2 = /(*)• (28-10) 



never vanishes inasmuch as y x and y 2 are linearly independent solutions of 
Eq. (28-3). 

The right-hand members of Eqs. (28-11) are known functions of x } and 
on integrating them we obtain fi(x) and v 2 (x ). We can thus write an in- 
tegral of (28-1) in the form 

V *= y i(z) [ ——dz + y 2 {x) f ~ ~ — ~ dx , 

J W(yi,y 2 ) J W{y u y 2 ) 

obtained by inserting v x (x) and v 2 (x) in (28-4). 


(28-12) 



74 


ORDINARY DIFFERENTIAL EQUATIONS 


(chap. 1 


Example 1. Find an integral of 


a?y ' — 2xy' 4-2 y « x log x, 


if x > 0. 


(28-13) 


It is easily checked that a pair of linearly independent solutions of the homogeneous 
equation associated with (28-13) is y\ « i, 1/2 « x 2 . Thus its general solution is y » 
C\x 4“ c*x 2 . Accordingly, we seek an integral of (28-13) in the form 


y ** v\x 4- t^ax 2 . 


(28-14) 


On dividing (28-13) through by x 2 to roduce it to the standard form (28-1), we see that 
f{x) « (log x)/x. Thus, Eqs. (28-6) and (28-10) yield 


VjX 4 - V 2 X 2 

tql 4“ v 2 2x 

Solving these for v[ and v 2 we obtain 

logx 

“ j 

x 


0, 

logx 


Vl m 


V 2 


and thus 


Pl ”-/ 


log! 


dx, 


log* 
~2 > 


' logx 


e ?2 * j dx. 


Integrating these and dropping integration constants (for any integral will do), we find 


Vi » -3^(logx) 2 , 


V 2 


— -(14- logx). 
X 


( 28 - 15 ) 


The substitution from (28-15) in (28-14) yields the desired integral of (28-13) in the form 
V » -x{l 4- logx 4- TsKlog x) 2 J 

Of course, we could have obtained this result directly from formula (28-12). 

The foregoing procedure can be generalized to compute an integral of 

y {n) + Pi(z)v in ~ 1) 4 b Vn~\{x)y f -f p(x)y = f(x). (28-10) 

If y i(x),. . ,, y n (x) is a known set of linearly independent solutions of the 
corresponding homogeneous equation, we seek an integral of (28-10) in the 
form 

V * v x (x)yi(x) + v 2 (x)y 2 (x) 4 b v n (x)y n (x), (28-17) 

where the v % (x) are unknown functions. To determine them we form the 
set of n — 1 equations by equating to zero the terms involving the v[ (x) 
in the expressions resulting from differentiating (28-17) successively n — 1 
times. The nth equation is got by inserting the corresponding values of 
derivatives in (28-10). We illustrate the procedure by an example. 1 

1 See also Prob. 5 at the end of this section. 



SEC. 28 ] LINEAR DIFFERENTIAL EQUATIONS 75 

Example 2. Find an integral of 

111 

v'" + y' ~ ^ V “ - t log *» * 1 * o. (28-18) 

A set of linearly independent solutions of the corresponding homogeneous equation is 
known to be 1 

yi - V 2 x log X , y$ « x(log x) 2 . (28*19) 

Accordingly, we take the integral of (28-18) in the form 

y •* v ix 4 log x 4 - V 3 x(Iog a:) 2 . (28-20) 

For the third-order equation the procedure just sketched yields the system of three equa- 
tions: 

V\Vx 4 HV% 4- v$ys « 0 , 

v’lVi 4 t» 2?/2 4 0 , (28-21) 

viyi 4t^2 +W «/(*). 

The reader will verify that, on setting /(x) « ( 1 /x 2 ) logx and noting (28-19), the sys- 
tem (28-21) yields 

» i- (log x) 3 , t >2 ~ (log x) 2 , ** - — log x, 

and we can take 

t>i - Hflogz) 4 , *2 - -H(iogx)*, vs - Kflogx ) 2 
Substitution in (28-20) gives finally y * (x/24)(log x)*. 

PROBLEMS 

1 . Use the method of variation of parameters to find integrals of the following equa- 
tions with constant coefficients, 

(а) y' 4 3 y ^ x 3 ; (c) y" ~ 2 y' 4 y » x; 

( б ) /' 4- V 4 by «■ e x ; (d) y ,n - 3 y' 4 2 y « 2(sin x - 2 cos x). 

2. Find the solution of 

^ 4 f\(x)y « / 2 (x) 

by the method of variation of parameters, and compare your result with that of Sec. 
10. The solution of the related homogeneous equation is obtained easily by separation 
of the variables 

8 . By the method of variation of parameters, find a particular integral of 
d 2 y Sdy 5 

~ y a log x, 

dx 2 x dx x 2 

whore the general solution of the related homogeneous equation is 

Cl . b 
V - — 4 c 2 x & . 
x 


4 See Example 1, Sec. 30. 



70 


ORDINARY DIFFERENTIAL EQUATIONS 


[chap. I 


4 . find the general solution of 

d*y x dy _ 1 

dx* 1 — xdx 1 — x ^ 


- x > 


where the general solution of the related homogeneous equation is eje* 4- e*x. 

6. Show that the formula corresponding to (28-12) for an integral of (28-16) is 


y(x) 


' X) V>{x) J 


w i(y h yt,. . ,,y n ) 
~W(y h yt,.. .,y n ) 


f(x) dx, 


where W(yi t p 2 >* . ,,y n ) is the Wronskian and Wi is the determinant obtained from W 
by replacing the ith column by (0,0,0,, . .,1), 


29. Reduction of the Order of Linear Equations. The method of vari- 
ation of parameters can be used to reduce the solution of every nth-order 
linear homogeneous equation to the solution of a linear equation of order 
n — 1 when one solution of the nth-order equation is known. This matter 
is of some importance in deducing general solutions of second-order linear 
equations, because one integral of such equations can often be determined 
by inspection. 

Let yi(x) be a solution of 


y" + Pi 0)2/' + p 2 (x)y = 0, (29-1) 


bo that y = cyi(x) is a solution for any constant c. If we replace c by an 
unknown function v(x) and seek a solution of (29-1) in the form 


y = (2&-2) 

we get, on differentiating (29-2), 

y f * vy'i + v'y u (29-3) 

y" « vyl + 2v f y[ + v"y x . 


Substituting from (29-2) and (29-3) in (29-1) and noting that y\{x) is a 
solution of (29-1), we get a separable equation 


v”yx 


for v(x ). 

Separation of variables in 


+ v'( 2 y\ + PiVi) « 0 
(29-4) gives 


so that 




y i 


-2 log yi-fpi dx. 


(29-4) 


Hence 


y? 


e~f p > ix . 


We see that v'(x) 9 * 0, so that v 9 * const. 


(29-6) 



LINEAR DIFFERENTIAL EQUATIONS 


77 


sec. 29] 

Integrating (29*5) we obtain 

v » J yr Z e~~f p ' dx dx, (29*6) 

so that the second linearly independent solution of (29-1), by (29*2), is 
V2 * Vx (x) jyi 2 (x)e^S p ^ ) dx dx. (29*7) 

We dispense with quite analogous calculations showing that the solution 
of an nth-order linear equation can be reduced to that of a linear equation 
of order n — 1 when one integral of the nth-order equation is known. 

Example * The equation 


with x 7 * Y% has an obvious solution vi 253 x. To determine another solution we set 
y «* vx. The function v f determined by formula (29-6), is 


v 


*-<)*' dr 

J x ~2 ( /2r-\ lag (2x— 1 ) 

I^'-J 


dx 


x 





dx 



X 

Thus the second solution is y * vx ~ e 2x . 


PROBLEMS 


1, The equation x 2 y" + 2 xy' ** 0 has an obvious solution 
is another solution, and thus find the general solution. 

2 . One solution of 


y n + : 


- 2x 


X 2 + X 


2x 2 - 2x_ 
j 3 -f x 


V 


y 


1 . Show that y 


0 


l/x 


obviously is y » x 2 , Show that a second solution is y xc~~*. 

3. A special case of Ixjgendre’s equation 

(1 - x % )y" ** 2 xy' + 2y - 0 

has an obvious solution y *» x. Obtain a first-order equation for a second linearly 
independent solution of this equation, and solve. 



78 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 


SO. The Euler-Cauchy Equation. An equation of the form 

x*V (B) + ai* n ~V B '~ 1) + • • • + a n ^xy' + a n y = /(*), (30-1) 

where the a t are constants, is usually called Cauchy’s equation, although 
it was examined earlier by Euler. 'We show that by a change of the in- 
dependent variable x, it can be transformed into an equation with constant 
coefficients which can be solved by familiar methods. 

If we set 

x = e*, ( 30 - 2 ) 

dx dz 

then — » e t and — e”**. 

dz dx 


On writing D « d/dz, we get 


y f 



dy dz 
dz dx 


e g Dy. 


Also 


dy' dy' dz 
dx dz dx 


W)e- 


= e~ 2, (Z) 2 - D)y 
- e~ 2 ‘D(D - 1 )y. 

In a similar way we find 

V (n> = c- n ‘D(D - 1 ){D — 2) • • • (D - n+ l)j/. (30-3) 

From (30-2), x n = e"*, and the substitution from (30-3) in (30-1) there- 
fore yields the equation with constant coefficients 


\D(D - 1 ){D — 2) •••(£> — n + 1) 


+ aiD(D — 1) ••• (£> — « + 2) -f 1- 4- a n ]y — /(«'). (30-4) 

If a solution of (30-4) is denoted by y = F(e), then the solution of the 
original equation, as follows from (30-2), is y = F( log x). 

Example 1. Find the general solution of 

*V" + *»'-»-* log I. (30-6) 

Upon setting x «* «\ this equation becomes 

0>0> - 1)0> - 2) + ^ 

or (D® — 3D 2 -f 3/) — l)y » ze*. (30-6) 

The roots of the characteristic equation obviously are mi ** m% ** m* ■» 1 . Hence the 
solution of the homogeneous equation is (cj -f caz + c»z 2 )e*. 



79 


SEC. 31] APPLICATIONS OP UNEAR EQUATIONS 

Inasmuch as the characteristic equation has a triple root and the right-hand member 
of (30-6) is a solution of the homogeneous equation, we take the trial integral in the form 
(see Sec. 27) 

y * azV. 

The substitution of the trial integral in (30-6) shows that a *■ so that the general 
solution of (30-6) is 

y ■* (a 4 C 2 z 4 c&~)e* -f Kaz***. 

Finally, the substitution of z » log x gives 

y « ki 4 <*2 log x 4 r 3 (log x) 2 jx 4 H 4 j(log x) 4 , 
which is the desired solution of (30-5). 

The general solution of the homogeneous equation 

xV n) + ai* n ~'!j (n ~' ) -f h a ri ,.ixy' + a n y = 0, 

associated with (30-1), can often he found by taking a trial solution y = x m . 
This is illustrated in the following example. 

Example 2. Solve x 2 y n + 2x?/ * 0 The substitution of y ~ x m yields the equation 
m(m — l)x w 4 2 wx m =* 0, 
or m(w — 1) 4- 2m -= 0. 

Since m * 0 and m » —1 satisfy this equation, y =* x° ~ 1 and y - x ~ l are linearly 
independent solutions of the given equation The general solution, therefore s 

y ** ci -f C2X*' 1 . 


PROBLEMS 

Find the general solutions of: 

1. x 2 //" 4- 4xi/' 4 2iy « log x; 2. xV" ~ 4x 2 </" 4* 5 xy r — 2y *■ 1; 

3. jt 7 y” 4 y ~ x 2 ; 4. x 2 y " - 2xy * 4 2y - x logx. 

By assuming a solution of the form y = x m solve: 

6. xV' — 4xy' 4 6?/ «* 0; 6. x 2 y" 4 2xy' - n(r> 4 l)y ~ 0. 


APPLICATIONS OF LINEAR EQUATIONS 


31. Free Vibrations of Electrical and Mechanical Systems. 

Sec. 19 that the equation 


d(mv) 

dt 


F(SyV,t), 


We saw in 


(31-1) 


stating Newton's second law of motion, is readily integrable when the 
external force F is a function of the displacement s alone, when it is a 
function of the velocity v alone, or when it depends only on the time t. 



80 ORDINARY DIFFERENTIAL EQUATIONS (CHAF. 1 

In this section we examine other types of this equation which are of cardinal 
importance in the analysis of oscillating electrical and mechanical systems. 

Throughout this discussion we shall assume that the mass m is constant, 
so that Eq. (31-1) can he written as 

d 2 s 

m '^2 == F( - S > V ’^> 

where v m ds/dt. 

We begin our study with a simple mechanical system which is a proto- 
type of more general systems that appear in the analysis of vibrations of 
elastic structures. 

Let it be required to determine the position of the end of an elastic 
spring set, oscillating in a vacuum. If a mass M is applied to one end of 
the spring whose other end is fixed, it will produce the elongation $, which, 
according to Hooke’s law, is proportional to the 
applied force F = Mg, g being the gravitational 
acceleration. Thus, 

F — ks = Mg , 

where k is the stiffness of the spring. 

If at any later time t an additional force is ap- 
plied to produce an extension //, after which this 
additional force is removed, the spring will start 
oscillating. The problem is 1o determine the posi- 
tion of the end point of the spring at any subse- 
quent time. 

The forces acting on the mass M are the force of 
gravity Mg downward, which will be taken as the 
positive direction for the displacement y f and the 
tension T in the spting, which acts in the direction 
opposite to that of the force of gravity (Fig 10). Hence, from Newton's 

second law of motion, 0 

d y 

M ~ « Mg - T. 
dr 

Since T is the tension in the spring when its elongation is 8 + y, Hooke's 
law states that T « k(s + y), so that 

d 2 y 

M~ « Mg - k(s + y). 

dr 

But Mg « ks, and therefore the foregoing equation becomes 

d 2 y 

M — + ky^ 0. 
dr 





SEC, 31] APPLICATIONS OF LINEAR EQUATIONS 

Setting k/M *» a 2 reduces this to 

~ + a 2 y « 0 or (D 2 + a 2 )y » 0. 

dr 


81 


(31-2) 


Factoring gives (I) 
general solution is 

or, in real form, 


at)(D + ai)y - 0, from which it is clear that the 
y = Cie~ alt + c 2 c alt , 
y ~ A cos at + B sin at 


The arbitrary constants i and i? can be determined from the initial 
conditions. The solution reveals the fact that the spring vibrates with a 
simple harmonic motion whose period is 



"oifv 


The period depends on the stiffness of the spring as would be expected — 
the stiffer the spring, the greater the frequency of vibration. 

It is instructive to compare the solution just obtained with that of the 
corresponding electrical problem. It will be seen that a striking analogy 
exists between the mechanical and electrical 
systems. This analogy permits one to replace 
a study of complicated mechanical systems by 
the analysis of performance of mathematically 
equivalent simple electrical circuits. 

Let a condenser (Fig. 17) be discharged 
through an inductive coil of negligible resist- 
ance. It is known that the charge Q on a con- 
denser plate is proportional to the potential difference of the plates; that is, 

Q = CF, 


L 

Fro. 17 


where C is the capacity of the condenser. Moreover, the current I flowing 
through the coil is 



and, if the inductance be denoted by L, the emf opposing V is L dl/dt , 
since the IR drop is assumed to be negligible. Thus, 



82 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

Simplifying gives 

7? + n Q -°’ 

which is of precisely the same form as (31-2), where a 2 « 1/CL, and the 
general solution is then 

t t 

Q = A cos — 7 == + B sin — r— * 

VCL VCL 

The period of oscillation is 

T = 2irVcL. 


Note that we can make the inductance L correspond to the mass M of the 
mechanical example and 1/C correspond to the stiffness k of the spring 

32. Viscous Damping. Let the spring of the mechanical example of 
Sec. 31 be placed in a resisting medium in which the damping force is 
proportional to the velocity. This kind of damping is termed viscous damp- 
ing. 

Since the resisting medium opposes the displacement, the damping force 
r(dy/dt ) acts in the direction opposite to that of the displacement of the 
mass M . The force equation, in this case, becomes 


or, since Mg * ks , 


d 2 y 

M TJ ~ Mg 
dr 


<i 2 y 


k{y + «) - r 


Ay 

dt 


r dy k 

dt 2 + M~dt + M y ~ °' 


To solve this equation we write it in the more convenient form 


d 2 y 

dt 2 


+ 26 — + a 2 y 
at 


0 , 


(32-1) 


where 26 = r/M and a 2 — k/M. In this case the characteristic equation 
is 

m 2 + 26m + a 2 = 0 

and its roots are 

m = — 6 ± vV - o 2 , 
so that the general solution is 

y * )t + ^(-b-s/b'-a* )t (32-2) 

It will be instructive to interpret the physical significance of the solution 
(32-2) corresponding to the three distinct cases that arise when b 2 — a 2 > 0, 



SEC. 32) APPLICATIONS OF LINEAR EQUATIONS 83 

b 2 — a 2 = 0, and b 2 ~~ a 2 < 0. If 6 2 — a 2 is positive, the roots m are 
real and distinct. Denote them by m x and m 2y so that (32-2) is 

y ~ (32-3) 

The arbitrary constants c*i and c 2 are determined from the initial conditions. 
Thus, let the spring be stretched so that y = d and then released without 
giving the mass M an initial velocity. The conditions are then 


when t ® 0 and 


y ~ d 


dy 

dt 


- 0 


when t == 0. 

Substituting these values into (32-3) and the derivative of (32-3) gives 
the two equations 


d « C\ + c>> and 0 = mic x + w 2 c 2 . 
These determine 

m 2 d mid 

c x * and c 2 = 

nil *" TWj — w 2 


Hence, the solution (32-3) is 

d 

y a- (m x e mit — m 2 e m ' t ). 

nil ^2 


The graph of the displacement represented as a function of t is of the type 
shown in Fig. 18. Theoretically, ?/ never becomes zero, although it comes 


arbitrarily close to it. This is the 
so-called overdamped case. The re- 
tarding force is so great in this ease 
that no vibration can occur. 

If b 2 — a 2 = 0, the two roots of 
the characteristic equation are equal 
and the general solution of (32-1) be- 
comes 

y = + c 2 t). 



If the initial conditions are y * d > dy/dt * 0 when £ = 0, the solution is 

y = + 60- 


This type of motion of the spring is called deadbeat . If the retarding force 
is decreased by an arbitrarily small amount, the motion will become 
oscillatory. 



ORDINARY DIFFERENTIAL EQUATIONS 


84 


[chap* 1 


The most interesting case occurs when b 2 < a 2 , so that the roots of the 
characteristic equation are imaginary* Denote b 2 — a 2 by —a 2 , so that 


m 


-b dtz ia 


and 


y « 




+ c 2 fi 


(— 6—icr)( 


e *%4 cos at + B sin at). 


If the initial conditions are chosen as before, 

y = d 
dy 


when t ** 0 and 




- 0 


when * 0, the arbitrary constants A and B can be evaluated. The result 


is 


y * de bt ^cos at H — sin * 

which can be put in a more convenient form by the use of the identity 

A cos 6 + B sin 6 s= a/a 2 + # 2 cos ^6 ~ tan” 1 — ^ • 

The solution then appears as 

y - - VaM- b 2 e- bi cos (orf - tan" 1 ~Y (32-^4) 

a \ a/ 


The nature of the motion as described by (32-1) is seen from Fig. 19. 
It is an oscillatory motion with the amplitude decreasing exponentially. 

The period of the motion is T = 2ir/a. 



An electrical problem corresponding to the example of the viscous 
damping of & spring is the following: A condenser (Fig. 20) of capacity C 
is discharged through an inductive coil whose resistance is not negligible* 



SEC. 32 } APPLICATIONS OP LINEAR EQUATIONS 86 

Referring to Sec. 31 and remembering that the IR drop is not negligible, 
we find the voltage equation to be 


or 

Simplifying gives 


V 


L 7/2=0 

(it 


Q d 2 Q dQ 
- + Z/-— + /2 — 
C dt 2 dt 

d 2 Q R dQ Q 

~dt i + L It + CL 


= 0. 
= 0, 


1 — os 

i 


and this equation is of the same form as that in the mechanical example* 
The mass M corresponds to the induct- 
ance Lj r corresponds to the electrical 
resistance R } and the stiffness k cor- 
responds to 1/C. Its solution is the 
same as that of t lie corresponding me- 
chanical example and is obtained by 
setting 2b = R/L and a 2 ~ 1/CL. p IG 20 


-VWW 1 

R 


PROBLEMS 


1. The force of 1,000 dynes will stretch a spring 1 cm A mas* of 100 g is suspended 
at the end of the spring and set vibrating. Find the equation of motion and the frequency 
of vibration if the mass is pulled down 2 cm and then released What a ill be the solution 
if the mass is projected down from rest with a velocity of 10 cm per sec? 

2. Two equal masses are suspended at the end of an elastic spring of stiffness k. One 
mass bills off Describe the motion of the remaining mass. 

3. The force of 98,000 dynes extends a spring 2 cm. A mass of 200 g is suspended at 
the end, and the spring is pulled down 10 cm and released Find the position of the 
mass at any mat ant / if the resistance of the medium is neglected. 

4 . Solve Prob 3 under the assumption that the spring is viscously damped. It is 
given that the resistance is 2,000 dynes for a velocity of 1 cm per sec. What must the 
resistance be in order that the motion be a deadbeat? 

6. A condenser of capacity 4 is charged so that the potential difference of the plates 
is 100 volts The condenser is then discharged through a cod of Resistance 500 ohms 
and inductance 0 5 henry. Find the potential difference at any later time i. How large 
must the resistance be in order that the discharge just fails to be oscillatory? Determine 
the potential difference for this case. Note that the equation in this case is 



r dV V 

+ , ‘li + c 


0. 


6. Solve Prob. 5 if R * 100 ohms, C * 0.5 nt, and L « 0 001 henry. 

7. A simple pendulum of length l is oscillating through a small angle 9 in a medium 
in which the resistance is proportional to the velocity. Show that the differential equa- 


tion of the motion is 


<Pe 

dt 1 


do g 

+2k h + j 6 


Q. 


Discuss the motion, and show that the period is 2*7 Vw 2 — k? where w 2 m g/l. 



ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 

8. An iceboat weighing 600 lb is driven by a wind that exerts a force of 25 lb. Five 
pounds of this force is expended in overcoming frictional resistance. What speed will 
this boat acquire at the end of 30 sec if it starts from rest? Bint: The force producing 
the motion is F » 25 — 5 * 20. Hence, 500 dv/dt * 20 g. 

9. A body is set sliding down an inclined plane w r ith an initial velocity of t\> fps. If 
the angle made by the plane with the horizontal is 0 and the coefficient of friction is p r 
show that the distance traveled in t sec is 


8 ** sin 0 — fx cob 0)t 2 4* 


Hint: m dPs/dt 2 * mg sin 0 ~ pmg cos 0. 

10. One end of an elastic rubber band is fastened at a point P t and the other end 
supports a mass of 10 lb. When the mass is suspended freely, its weight doubles the 
length of the band. If the original length of the band is 1 ft and the weight is dropped 
from the point P, how far will the band extend? What is the equation of motion? 

11. It is shown in books on strength of materials and elasticity that the deflection of 
a long beam lying on an elastic base, the reaction of which is proportional to the deflec- 
tion y, satisfies the differential equation 


El 


<?y 

d? 


~ky 


8et o 4 «■ k/4EI, and show that the characteristic equation corresponding to the result- 
ing differential equation is m A 4 4a 4 — 0, whose roots are m «* db a dr at. Thus show 
that the general solution is 


y — cie ax cos ax 4 C 2 e ax sin ax 4 W”** cos ax 4 c\e~~ ax sin ax. 


12. If a long column is subjected to an axial load P and the assumption that the curva- 
ture is small is not made, the Berrioulh-Euler law gives (see Sec. 5) 


P 



y 

Fig. 21 


c Py/dx 2 _ M 

(1 4 (dyldxW m W 

Since the moment Af is equal to - Py (Fig. 21), it follows upon setting 
dy/dx » p that the differential equation of the deformed central axis is 

yjdp/dy) ^ _ Py 
(1 4 V) h ^ El 

Solve this differential equation for p , and show that the length of the 
central line is given by the formula 



where k 2 «* d 2 P/4E/, d is the maximum deflection, and F(k,v/2) is the elliptic integral 
of the first kind [see Eq. (20-12)]. The equation of the elastic curve, in this ease, cannot 
be expressed in terms of elementary functions, for the formula for y leads to an elliptic 
integral See, however, Chap. 2, Sec. 10. 

S3. Forced Vibrations. Resonance. In the discussion of Sec. 32, it was 
supposed that the vibrations were free. Thus, in the case of the mechanical 
example, it was assumed that the point of support of the spring was station- 



SEC. 33 ] APPLICATIONS OF LINEAR EQUATIONS 87 

ary and, in the electrical example, that there was no source of eraf placed 
in series with the coil. 

Now, suppose that the point of support of the spring is vibrating in 
accordance with some law which gives the displacement of the top of the 
spring as a function of the time t, say x — /(<), where x is measured positively 
downward. Just as before, the spring is supposed to be supporting a 
mass M, which produces an elongation 8 of the spring. If the displace- 
ment of the mass M from its position of rest is y, it is clear that when the 
top of the spring is displaced through a distance x t the actual extension 
of the spring is y ~~ x. If the resistance of the medium is neglected, the 
force equation is 

d 2 y 

M — ~ Mg - k(s + y - t) » -~k(y - x), 
dr 

whereas if the spring is viscously damped, it is 

d 2 y dy 

dt 2 di 


Upon simplification this last equation becomes 


M 


d 2 y 

dt 2 


dy 

+ r — + ky ~ kx } 
dt 


( 33 - 1 ) 


where x is supposed to be a known function of t. 

The corresponding electrical example is that of a condenser (Fig. 22) 



which is placed in series with the source of emf and discharges through a coil 
containing inductance and resistance. The voltage equation is 

dl 

-RI-L-+V -/(*), 
dt 

where /(/) is the impressed emf given as a function of L Since 



88 


ORDINARY DIFFERENTIAL EQUATIONS 


the equation becomes 


d 2 V dV 

cl-~ t + cr--+ V 

dt 2 dt 


■m- 


[chap. 1 
(33-2) 


An interesting case arises when the impressed emf is sinusoidal; for example, 

f(l) = Eo sin wl. 


Then the equation takes the form 


d 2 V RdV 1 
'd J + L dt + CL 


1 

V = — E o sin wt. 


Both (33-1) and (33-2) are nonhomogeneous linear equations with 
constant coefficients of the type 

+ 2& •— + an/ = <rf(t). (.53-3) 

dt~ at 

The solution of this equation is the sum of the complementary function 
and a particular integral. The complement ary function has the form 
(32-2), namely, 

CxC m d + Co'’ 7 " 2 *, 

where 

m x = — b + vV — a 2 and rn 2 ~ -b — \/b 2 — a 2 . 

A particular integral y = u(t) can be deduced for Eq. (33-3) for an arbi- 
trary continuous function /(/) bv the method of variation of parameters 1 
If the impressed force /(/) in (33-3) is simple harmonic of period 2 v, w 
and amplitude a [h then 

/(/) = a 0 sin wl 

and an integral y — u{t) can be obtained by the method of Sec. 25. The 
result is 

a 2 ao 

yv> - v<? +«v * ( “‘ - <>- (33 - 4) 

—1 

where « = tan ~* 

a — w l 

From discussion in See. 32, it is clear that the part of the general solution 
of (33-3) which is due to free vibrations is a decreasing function of t, 
becoming negligibly small after a sufficient lapse of time. Thus, the 1 'steady- 
state solution” is given by the particular integral (33-4). 

* flee the corresponding computation at the end of this section for the case when b ** 0. 



SEC. 33] APPLICATIONS OP LINEAR EQUATIONS 89 

It must be observed that when the impressed frequency u> is very high, 
the amplitude of the sinusoid in (33-4) is small. When w is nearly equal 
to the natural frequency a, the amplitude is nearly a 0 a/2b. This may be 
dangerously large if the resistance parameter b is small. For a and b 
fixed, the maximum amplitude occurs when 

~ ((a 2 — w 2 ) 2 + 46 2 w 2 ] — 0, 
do) 

that is, when 

O. 2 = a 2 - 2b 2 . (33-5) 

Stated in terms of the physical quantities of electrical and mechanical 
examples, a large amplitude in (33-4) means a large maximum emf, or a 
large maximum displacement of the spring. These, as we have already 
noted, may become excessively large when the resistance r of the medium is 
small and the impressed frequency w is close to the natural frequency a. 
This phenomenon, known as resonance , is of profound importance in 
numerous engineering and physical situations. 1 
If b — 0, Eq. (33-3) reduces to 

d 2 y 

- - + a 2 y = a 2 f{t). (33-6) 

at 

We can easily deduce a formula for an integral y(t) of (33-6) for an rrbitrary 
forcing function f(t). Since sin at and cos at are linearly independent 
solutions of Eq. (33-6) with f(t) — 0, the method of variation of param- 
eters of Sec. 28 suggests taking a solution in the form 

?/(0 ~ ?>i (/) cos at T v 2 (l) sin at (33-7) 

For the determination of v\ (/) and v 2 (f) we have a pair of equations [see 
Eqs. (28-6) and (28-10)] 

v\ cos at + v 2 sin at ~ 0, 

— av\ sin at + av 2 cos at = a 2 /(f). 

Solving these for v\ and v 2 we get 

v[ == ~af(t) sin at, v 2 = af(t) cos at, 
which on integration between the limits 0 and i yield 

vM) =* —a[ f{t) sinatdt, v 2 (t) ~ af f(t) cos at dt 
J 0 ■'0 

1 The failure of the Tacoma bridge was explained by some authorities on the basis of 
resonant forced vibrations, and there are instances of the collapse of buildings induced 

by the rhythmic swaying of dancing couples. The failure of propeller shafts is often 
attributed to forced torsional vibrations. Sec also Joshua 6:5. 



00 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

Formula (33-7) then yields 

y(t) « — a cos atj^f(\) sin a\d\ + a sin atf^f(\) cos aX dX, (33-8) 

in which we have replaced the integration variable t by X so as not to con- 
fuse it with the variable t in the limits. It follows directly from (33-8) 
that the integral y(t) corresponds to the initial conditions y( 0) * y'(Q) * 0. 
If we combine integrals in (33-8), we get the desired formula 

y{t) a* a f /(X) sin a(t — X) d\. (33-9) 

J o 

When the forcing function /(0 is taken in the form f(t) — a 0 sin at (so that 
the impressed frequency is equal to the natural frequency), this formula 
yields 

y{t) = aao f sin aX sin a(t — X) dX. 

J o 

After simple integration we obtain 


y = — (sin a£ — cos at ), 

2 


representing a vibration whose amplitude increases with time, for the 
amplitude a 0 /2 in the first term is constant and the amplitude of the 
second term, a 0 at/2, grows with t. In any physical situation, some resist- 
ance is present, and a reference to (33-4) shows that b prevents oscillations 
from becoming arbitrarily large. Nevertheless, they may be dangerously 
large if b is small and a is near w. 


PROBLEM 

Obtain a formula for a particular integral of Eq. (33-3) analogous to (33-9), and deduce 
from it the result (33-4). The integration will be simplified if sin a >t is replaced by 
(< e <wt - e~^)/2i. 

34. The Euler Column. Rotating Shaft It is known from experiments 
that a long rectilinear rod subjected to the action of axial compressive 
forces is compressed and retains its initial shape as long as the compressive 
forces do not exceed a certain critical value. Upon gradual increase of the 
compressive load P, a value of P — P% is reached when the rod buckles 
suddenly and becomes curved. The deflections of rods so compressed 
become extremely sensitive to minute changes of the load and increase 
rapidly with the increase in P. A detailed analysis of this instability or 
buckling phenomenon depends on rather delicate considerations in non- 
linear theory of elasticity. However, if the argument of Euler is followed, 



SEC. 34] APPLICATIONS OP LINEAR EQUATIONS 91 

it is possible to deduce the magnitude of the critical load P\ from linear 
differential equations governing small deflections of loaded rods. 

Thus, consider a rod of uniform cross section and length l, compressed 
by the forces P applied to its ends (Fig. 23). Initially this rod is straight, 



Fig. 23 


but after the critical load Pi is reached, it becomes curved, and we denote 
the deflections of its central line by y. 

It is known from the Bernoulli-Euler law (5-2) that for small deflections 

d 2 y __ M 

d a* “ eT 


where, in our case, the bending moment M - — Py . Thus 



d 2 y Py 

dh? = ~ W 


or 

y" + k?y - o, 

(34-1) 

where 

MS 

hi 

* 

(34-2) 

Equation (34-1) must be 

solved subject to the end conditions 



3 

V ' 

II 

© 

<< 

s 

il 

p 

(34-3) 


since the ends of the rod remain on the x axis. 

The boundary-value problem characterized by Eqs. (34-1) and (34-2) is 
quite different from the initial-value problems considered heretofore. In 
the initial-value problems we seek solutions of differential equations 
satisfying specified conditions at one point only, while in the boundary- 
value problem stated above the solution y must satisfy conditions (34-3) 
assigned at two points x = 0 and x = l It is not obvious that a solution 
of a differential equation satisfying specified conditions at two points 
exists in general. We shall see, however, that for suitable choices of the 
parameter k Eq, (34-1) does have solutions vanishing at the end points 
X 0, x *■ l. 



0 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

Now the general solution of (34-1) is 

y = ci cos kx + c 2 sin kx (34-4) 

and, on imposing the conditions (34-3), we get two equations 
0 = Ci cos W) + c 2 sin kO 
0 = Ci cos kl + c 2 sin kl. 

rhese demand that 

ci * 0, c 2 sin kl — 0. (34-5) 


The choice c x c 2 = 0 gives y — 0, corresponding to the rectilinear shape 
of the rod. If the rod does not remain straight, c 2 n* 0, and we conclude 
from (34-5) that sin kl = 0, so that 


7?7T 

fc = y> n = 0,1,2,.... (34-6) 

The choice of n *= 0 again gives y ~ 0. If n = 1, k = 7 r/Z, and on recalling 
the definition (34-2), we see that the corresponding value of P is 

Pi (34-7) 

r 

This is the critical , or the Euler , food. 

The shape of the central line of the rod, in this case, is 


TTX 

y ~ c 2 siny* 

The choice of n = 2, 3, ... in (34-6) gives other “critical loads’' P 2 , /V 
. . . and the corresponding solutions 

m tz 

y - 02 sm ~ * 

The maximum deflection c 2 is not determined in this analysis, and, indeed, 
no far-reaching conclusions should be made from such calculations inas- 
much as they are based on the assumption of small deflections implicit 
in our use of the Bemouili-Euler law. 

Another interesting problem, essentially of the same sort, arises in the 
study of rotating shafts. It has been noted that when a long shaft sup- 
ported by bearings at x = 0 and x = l is allowed to rotate, its initially 
rectilinear shape is preserved only if the speed of rotation u> does not 
exceed a certain critical value u>\. On approaching the speed o>x the shaft 
starts pulsating and its shape changes. On further increase of the speed 
another critical value « 2 is reached when the shaft starts beating and its 



SEC. 84] APPLICATIONS OF LINEAR EQUATIONS 93 

shape changes again, and so on. This phenomenon can, in part, be explained 
by calculations similar to those used in determining the Euler load. 

Let us suppose that the shaft is rotating with the angular speed u>. An 
element of length dx of the shaft experiences the centrifugal force 

F dx = pdx u 2 y } 

where p is the density per unit length of the shaft and y is the deflection 
at the point x. Thus, 

F « pa> 2 y (34-8) 


is the force per unit length of the shaft distributed along its length. It 
is shown in books on strength of materials that when the forces F acting on 
a rod are normal to its axis, then 


F « 


d 2 M 

dx 2 


where the bending moment M is given by the Bemoulli-Euler law 


M - El 


d 2 y 

dx 2 ' 


(34-9) 


Thus, 



(34-10) 


and if the flexural rigidity El is constant, Eq. (34-10) reads 

d 4 y ^ F 

dx 1 ~ In 


(34-11) 


The substitution for F from (34-8) gives the desired equation for the 
rotating shaft: 

d 4 y 

-4 - Pjr = 0 (34-12) 

V* * 


dx 4 


with 


pet) 

El 


(34-13) 


Since the roots of the characteristic equation m 4 — A; 4 = 0 are rn = ifc, 
m » db kij the general solution of (34-12) is 

y = cie kx e 2 e" kx + c 3 cos kx + c 4 sin kx . (34-14) 

If at the points of support x ~ 0, x = l the deflection y and the moment 
ill are zero, then [see (34-9)] 

y( 0) - 0, y"(0) - 0, 

2/(0 - 0 , 


2/"(0) - 0, 
2/"(0 - 0 . 


(34-15) 



94 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 


The substitution from (34-14) into the boundary conditions (34-15) 
yields four equations: 

Ci + C 2 + C8 *= 0, 


d + C2 — C3 = 0, 

c x e kl + + c 3 cos kl + c 4 sin Id = 0, 

c X € hl + c 2 e~ hl — C3 cos kl — c 4 sin kl — 0. 


(34-16) 


The solution c x ~ c 2 = c 3 * c 4 « 0, yielding y - 0, corresponds to the 
straight shaft. The system (34-16) also has nonzero solutions for certain 
values of k. From the first two equations (34-16) we find 

c x ~ ~c 2 , c 3 — 0, 

and the substitution of these values in the two remaining equations gives 
C\ — c, 2 = c 3 = 0, c 4 sin kl = 0. 

Thus, sin kl * 0 unless c 4 = 0, and hence 


k — — » n = 1 , 2 , 

Z 

Using the value of k for n ® 1 in (34-13) gives the first critical speed 

t 2 jm 

Wl = J \T’ 

The critical speeds w 2 , c*>3, . . . are determined by taking k with n = 2, 3, 


PROBLEMS 


1 . When a beam lies on an elastic foundation, then in addition to the transverse ex- 
ternal load F(x) t there is a restoring force R — — a 2 y proportional to the deflection y . 
The equation of the axis of the beam then has the form 

Ely™ + a'y - F(x). 

Solve this equation for F(x) — p, a constant, by assuming that the ends of the beam are 
hinged so that 

3/(0) - y"(0) « y(l) - i/"(0 - 0. 


2. The differential equation of the deflection y of the truss of a suspension bridge has 
the form 

"3 ?-« + »§ -»- 4 ’ 


where J5T ■» horizontal tension in cable under dead load 9 
k ** tension due to live load p 
J? *» Young's modulus 



SYSTEMS OF EQUATIONS 


sec. 35] 


93 


I ■* moment of inertia of cross section of truss about horizontal- axis of truss 
through center of gravity of section and perpendicular to direction of length 
of truss 

Solve this equation under the assumption that p — qh/H is a constant. 

3. The differential equation of the buckling of an elastically supported beam under an 
axial load P has the form 


d A y P d*y k 
dx* + El da* + EI V 


0 , 


where El is the flexural rigidity and k is the modulus of the foundation. Solve this 
equation. 


SYSTEMS OF EQUATIONS 


35. Reduction of Systems to a Single Equation. We saw in Sec. 13 that 
it may prove advantageous to reduce the solution of a second-order equation 
to the solution of a system of two equations of first order. Thus the 
dynamical equation considered in Sec. 31, 


dh 

dt* 


F(s,s',t) 


with s' se ds/dt, can be reduced to a system of two equations, 


-r = 

at 



by setting s' = v. 

In the same manner, the third-order equation 


« F(y,y',y",t ), (354) 

in which y f se dy/dt and y n as dPy/di t 2 , is reducible to a system of three 
first-order equations in x lf x 2 , x 3 defined by 


y = *u v' * x 2y y” = * 3 . 


With these definitions, Eq. (35-1) can be replaced by a system of three 
equations: 

dxi dx 2 dx z 

— * * 2 , — * — = E(xi,x 2 ,X3,0. (35-2) 


This procedure can be extended to nth-order equations. 

A reduction of the nth-order equation to a system of n first-order equa- 
tions is of some practical importance in numerical integration of equations 
on differential analyzers and electronic calculators. Such computing de- 



m 


ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. I 

vices are usually so designed that it is simpler to calculate n first derivatives 
than one derivative of order n. The reduction has also numerous advan- 
tages in theoretical considerations. 

Systems of differential equations appear naturally in problems involving 
dynamical systems with several degrees of freedom. Thus, the motion 
of a particle constrained to move on a surface can be described by two 
positional coordinates 1 (x,t/). These coordinates satisfy equations of the 

form 9 

d x / dx dy \ 


d It 2 
d 2 y 
dt 2 


( dx dy \ 
**•*'*■')■ 


This pair of second-order equations can be reduced to a system of four 
first-order equations. 

Alternatively, a system of n first-order equations can usually he re- 
duced to a single nth-order equation. A general discussion of this problem 
is involved, and we confine our remarks to systems of linear equations, 
because such systems commonly occur in applications. 

A system of n first-order equations 


dy i 
dt 
dy_2 
dt 


= flllVl + «12?/2 H f* nVn + /l(0» 


— 02l't/l + 0,22V 2 + * h 02nVn + /2W1 


( 35 - 3 ) 


dy* 

dt 


— Onll/l O n 2 I /2 — - — |— Cl nn y n -f* 


in which the a tJ and the f t (t) are continuous functions of /, is called linear . 
If the/ t (0 are all zero, the system is called homogeneous. 

The system (35-3) is linear because the solutions of the associated homogeneous system 
satisfy the linearity properties stated in Bee. 21. Thus if 

| 4 ”( 0 , .... V»\t) 

and y?\D, .... y„\t) 

are any two solutions of the homogeneous system, then the set of functions 
OJ/S 1 ’ + «wF\ cu4” + <-2V?\ .... erf}' + r 2 „«> 
is a solution of the homogeneous system for any choice of the constants c. 


1 If a particle moves on a sphere, for example, x and y may be taken as the latitude 
and longitude, respectively. 



SEC. 35] SYSTEMS OF EQUATIONS 9? 

Furthermore, it can be shown that the homogeneous system associated with (36-3) has 
a set of n solutions 


vi u . rf». . 

..,yi‘> 

first solution 

yf\ yf , - 

• •, y» 1 

second solution 

yi“\ yj w , • 

...yr. 

nth solution 

such that the determinant 

yi” 

rf* -- 

- y { n l) 



Vl 2 ’ 

-- 

■ y»> 

^0. 


Vl* 1 

vS* - 

- y» B) 1 



The general solution of the system (35-3) is then given by the set of n functions 

Vx “ cj y\ u -f c 2 yP H f* c n y\ n) -f u t {t), i ~ 1,2, . . n, (35-4) 

where yi ** «i(0, Vs *= «2(0» . ?/n * w n (0 is any solution of the nonhomogeneous 

ostein and the c t are arbitrary constants. The solution (35-4) is general in the sense 
that the rs can be always chosen so that there is a unique solution of the system (35-3) 
satisfying the arbitrarily prescribed initial conditions: 

Ul(k) ** Vl 0 , J/s(<o) ** //20, ...» ?/ndo) ==* Vn0- 

We indicate next how a system of first-order linear equations with con- 
stant coefficients can ordinarily be reduced to an equivalent single linear 
equation with constant coefficients whose order is equal to the number of 
equations in the system. 

Consider the system of two equations 


dx 

— + a,x + a 2 y = /i(0, 
at 

. (35-5) 

du 

— + hx + b 2 y = f 2 (t). 


We introduce the operator D == d/dt and write (35-5) as 
(D + ajx + a 2 y = /i(0, 

&i* + (D + h*)» = co- 
operating on the second equation in (35-G) with (l/iq)(Z) + oj), we get 


(35-6) 


(D + a,)r + ^ (i> + a,)(Z> + & 2 )y = ^ (D + ai)/ 2 (<). (35-7) 

6, 6i 

If we subtract the first equation in (35-6) from (35-7), we get, on multiply- 
ing through by b u 

(D + a,)(D + b 2 )y - b t a 2 y - (D + a{)f 2 (t) - hfi(t). (35-8) 



ORDINARY DIFFERENTIAL EQUATIONS 


[CHAP. 1 

This is a second-order linear differential equation with constant coefficients 
whose right-hand member is a known function. Hence its general solution 
y y(t) can readily be obtained. 

The characteristic equation for (35-8) is 

(m + ui)(t n + b%) — ** (35-9) 

and if its roots m = ra — m 2 are distinct, the general solution of (35-8) 
is 

y = Cie TO i* + + u(t), 

where u(t) is a particular integral of (35-8). If (35-8) has a double root 
m i = m 2 , the corresponding solution is 

y * Cie m ' ( + + u(t). 

Having obtained y T we can compute the solution for x, without further 
integration, by substituting y(t) in the second equation in (35-5). Thus 

X{t) - b 2 y(l) --]• 

The procedure for reduction of larger systems or for systems of equations 
of order higher than 1 is similar. 1 

Example 1. Consider 

dx 

~~ 4 2x - 2y « t f 
at 


dy 

dt 


- 4 y 


or ( D 4 2)x -2 y =* t, 

-3x 4 (/> 4 l)y * 

Operate on the second of these equations with 4 2) to obtain 
~(D 4 2)x 4 HID 4 2 )(U 4 1) y - WP 4 2)e‘, 
and add this result to the first equation. The result is 

H(D 4 2)(D 4 1 )y ~2y= H(D 4 2)e* 4 <, 

which simplifies to 

(Z> 2 4 3Z> - 4 )y - 3e* 4 3f. 

This equation can be solved for y as a function of t , and the result can be substituted id 
the second of the given equations to obtain x. 


1 See Example 2, p. 99. 



SYSTEMS OF EQUATIONS 


SEC* 35] SYSTEMS OF EQUATIONS 99 

Example 2, Let the two masses M\ and M% be suspended from two springs, as indi- 
cated in Fig. 24, and assume that the coefficients of stiffness of the springs are fa and 
fa, respectively. Denote the displacements of the masses from their positions of equilib- 
rium by x and y. Then it can be established that the following equations must hold: 


Af 2 


d 2 y 

d? 


-My - x), 




(fix 

Mi dfi - 


x) — k\x . 

These equations can be simplified to read 


By setting 


A 

Mi 



<Fy , 

fa 

fa 


dt 2 * 

' Mi 

V Mi 

t fix 

fa 

y 4 

fa 4 fa 

d? 

Mi 

Mi 

«* c 


h , 

- b\ 



M 2 ' 



0 , 


• 0. 


Mi 

M x 





the equations reduce to 


(D 2 -f b 2 )y - b 2 x 
—b 2 my 4 {D 2 4 a 2 4 b 2 m)x ■ 


0 , 

0. 


Operating on the second of these reduced equations with (1 /6 2 m)(Z) 2 4 b 2 ) and adding 
the result to the first of the equations give 

(D 2 4 b 2 )(D 2 4 a 2 4 b 2 m)x - b<mx - 0 

or [D 4 4- (a 2 4 b 2 4 b 2 m)D 2 4 a 2 b 2 ]x = 0. 

This is a fourth-order differential equation which can be solved for x as a function of t. 
It is readily checked that 

x ** A sin ( cot — 0 

is a solution, provided that w is suitably chosen. There will be two positive values of 
«*> which will satisfy the conditions. The motion of the spring is a combination of two 
simple harmonic motions of different frequencies. 


PROBLEMS 


Solve the systems: 
dy 


dx 

dt 

dx 

dt 


dt 

3x — 2y, 


dy dx 
dt ^ dt 


2 y, 


dy 

dt 

dy 


“2 x -y; 


2z' 


2 . 


dx 

a s 

(fix 

a* 


dy 
v ’ Tt 

d 2 y 
dt 2 


x; 


V , 


x; 


dx 

dt dt 

6. (D 4* 1)* 4 (2D 4 1 )y - (D - \)x 4 (D 4 1 )v - 1. 

7. Determine the solution in Example 1 satisfying y( 0) - 0, x(0) *» 0. 



100 ORDINARY DIFFERENTIAL EQUATIONS 

8. The equations of motion of a particle of mass m are 


[char. 1 


d 2 x 
1 di* 


X, 


d l y 

l dt* 


Y, 


dh 

1 di 35 


where x, y, z are the coordinates of the particle and X t Y, Z are the components of force 
in the directions of the x, y, and z axes, respectively. If the particle moves in the xy 
plane under a central atti active force, proportional to the distance of the particle from 
the origin, find the differential equations of motion of the particle. 

9. Find the equation of the path of a particle whose coordinates x and y satisfy the 
differential equations 


d x , U lly 

m -~r + He -~ 

dt 2 dt 


Ee % 




0 , 


where H, E, e , and m are constants. Assume that x - y -- dxRU = dy/dl « 0 when 
t * 0. This system of differential equations occurs in the determination of the ralio 
of the charge to the mass of an electron 

10. The currents / j and 1 2 in the two coupled circuits shown in Fig. 25 satisfy the 

following differential equations* 

*1 M U 2 


j A'VW' — i j 

§ C 
o c 

3c 

n 


*VWW V 1 


oj o 

^lo §^2 
So 




tPh d*h , p rf/2 , /* 
Jw ^ + ; ' 2 ,/F + K2 dT + ^ 

w «W* . , . p rfli . h 

M dl- +l '¥ +Rl dt + cl 


0 , 


Reduce the solution of this system to that 
Fig. 25 of a single fourth-oider differential equation. 

Solve the resulting equation under the as- 
sumption that the resistances Ry and R 2 an* negligible. 


36. Systems of Linear Equations with Constant Coefficients. We have 
indicated in the preceding section how a system of linear equations with 
constant coefficients can be solved by reducing the problem to the solution 
of one equation of higher order. In this section we sketch another mode 
of attack on the problem of solving the homogeneous system 


dy 1 
dt 


Gllth + 0122/2 + * * ’ + O-lnVny 


dy 2 

— - = a 2 i2/i + 0222/2 4 — * 4~ 02n2/», 

at 


(36-1) 


dVn 

dt 


0nl2/l + 0n22/2 4 + 0an2/n» 



SEC. 36] SYSTEMS OF EQUATIONS 101 

with constant coefficients. A third method, based on Laplace transform, 
is given in Appendix B. 

Let us seek our solution in the form 

Vi(t) ^ k 1 e u , y 2 (l) = k 2 e u , y n (t) = k n e u , (36-2) 

where the constants ki and X are to be determined so that Eqs. (36-1) are 
satisfied identically. 

The substitution from (36-2) in (36-1) yields 

\kie Xt = (a n ki + a n k 2 H h amfc n )e X/ , 

\k 2 C Xt = {o. 2 \k\ + <*22 ^2 +**’■+■ <l' 2 nkn)e U , 

\k n P « (jlnlkl “f* 0*2^2 -f- * * * &nnkn)C . 

On dividing each equation by c Kt and transposing all terms to one side, we 
get the system 

(ail ~~ X)*! + 012^2+ • • •+ (i\nk n = 0, 

021^1 + (022 “ X)/f 2 + • • • + «2 nkn ~ 0, 

(36-3) 

0nl^l + 0n2^2+ ' * * + (fl«n ““X)/f n = 0. 

This system is a system of linear homogeneous algebraic equations for 
the unknown ks . It has an obvious solution 

Ic i A 2 — — * * * kfi ■— 0 

corresponding to the trivial solution 

yi = th = * ' * = Vn « 0. 

Since we are interested in solutions (36-2) which are not all zero, we must 
seek values of the kts which are not all zero. Now, a system of Eqs. (36-3) 
will have such solutions for the k t if, and only if, its determinant 1 

0n X aj2 • • • 0i* 

021 022 ~~ X * * * 02n 

0*1 0n2 ’ * ’ 0nn X 

The equation D = 0 is called the characteristic equation for the system 
(36-1). On expanding the determinant, we see that (36-4) is an algebraic 
equation of degree n in X, and thus it has n real or complex roots: 

X “ Xj, X * X 2 , . . X = X n . 

1 Bee Appendix A. 


- 0. (36-4) 




102 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

If all these roots are distinct, then corresponding to each root X * X* 
there will be a solution of the form (36-2), namely, 

Vt(0 « k x e u \ y 2 (t) =» k 2 e u \ y n (t) - k n e Ut . (36-6) 

The constants hi in (36-5) must satisfy Eqs. (36-3) with X replaced by X». 

When Eq. (36-4) has multiple roots, the forms of solutions corresponding 
to multiple roots are more complicated. One solution corresponding to a 
multiple root X ~ X t surely has the form (36-5), but there will ^lso be 
solutions in the form of polynomials in t multiplied 1 by e Xtt . 

To clarify this discussion we consider a simple example. 

Example: Solve the system 

— * 2j/i -f 3y 2 , 

at 

(36-6) 

dy 2 0 , 

— * 2j/i + 2/2. 

We take a solution in the form 

2/i - kye Xt , i /2 - k 2 e Xt . (36-7) 

The characteristic equation (36-4) now reads 


On expanding it we get 



3 

1 - X 


0. 


X 2 - 3A - 4 - 0, 


the roots of which are Ai « — 1, X 2 ** 4. Thus, corresponding to the root Ai ® — 1 f we 
have a solution 

2/i » k\e~* t y 2 * kze”*. (36-8) 

To determine k\ and k 2 we form the system (36-3), 


(2 — A)*i Zk 2 “ 0, 

2k y + (1 - X)k 2 - 0 , 


(36-9) 


set A ** —1, and solve it for the k&. The result is 


ky * — k 2 * 

Thus, one of the ks can be chosen at will. If we take ky « a, we see from (36-8) that one 
solution of (36-6) is 

2/i oe""*, Vi - —ae~\ (36-10) 

the constant o being arbitrary. 

Another solution is obtained by taking A « 4. It has the form 

2/1 * k x e At , y 2 * fae" (36-11) 


1 Recall the corresponding situation in Bee. 26. 



103 


SEC. 36] SYSTEMS OF EQUATIONS f 

with ka determined by Eqe. (36-9) with X « 4. We find this time 

fa m Hfat 

so that, again, one of the ks can be chosen at will If we take fa ■» b, we obtain a solution 

j/i * fee 4 *, ya « ££fee 4 *. (36-12) 

From (35-4) it follows that the general solution of the system (36-6) is obtained by form- 
ing a linear combination of solut ions (36-12) and (36-10). We thus get the general solu- 
tion 

l/i * ae~ l + fee 4 *, 2/2 ■* -~ae~ < + %be u . 

This solution could have been obtained more easily by the method of Sec. 35. Thus, 
on writing the given system in the form 

(D - 2)yi - 32/ 2 - 0, 

(36-13) 

~2j/i + (D - 1 ) 2/2 - 0, 

we operate on the first equation with l /i(D — 1), add the result to the second, and get 
X(D - 1 )(D - 2)yi - 2y, - 0. (36-14) 

The corresponding characteristic equation is 

Him - 1 )(m - 2 ) - 2 - 0 , 

or m 2 — 3m — 4 *= 0. 


Since its roots are mi » — 1, m 2 ** 4, the general solution of Eq. (36-14) is 

V\ - -f c 2 € 4t . 

From the first of Eqs. (36-13) we have 

1/2 « H(D - 2)2/i » - 2)<c l e- < + <**“) * -c lC ~‘ + 

This checks the result found previously. 


The main object of this section is not so much to provide a new method 
for solving systems of linear equations but to introduce a few ideas on 
which the important study of stability of solutions of differential equations 
is based. There are several notions of stability of solutions, and we illus- 
trate only two such by considering some simple examples. 

The system 

dy 


dt 


* x > 


— = —2 bx — a 2 y t 


(36-15) 


is, obviously, equivalent to one second-order equation 


<fy 

di l 


+ 25 


dy 

dt 


+ a 2 y = 0 . 


(36-16) 



104 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

As we saw in Sec, 32, its general solution when b 2 — a 2 > 0 is 

y * c 1 e < - b+V,? ~ SS >* 4- c 2 e { - > '~ Vb ‘-° i)t . (36-17) 

If b 2 — a 2 < 0, we can write (36-17) as 

y = cos Va 2 — b 2 t + B sin vV — b 2 t ). (36-18) 

If b * a, we have the solution 

y = e~ 6 *(c L + c 2 0* (36-19) 

For 6 « 0, we have the equation 

— + a 2 y = 0, (36-20) 

whose general solution is 

y — A cos at + B sin at. (36-21) 

We observe that if b > 0, the solutions (36-17) and (36-19) are damped. 
That is, | y(£) | —* 0 as t — ► *». If b < 0, these solutions are not damped 
because \y(t) | — > °o for a sequence of \ allies t —> ^ As regards the case 
6 = 0 f we see from (36-21) that y{i) oscillates between -f V/l 2 + B 2 and 
— VA 2 + B? [see the iormula just above (32- 1)]. 

If we write Eq. (36-16) in the form 

d 2 y _ du 

— + a 2 i» = -2 by', y' = -- (36-22) 

dr dt 

and compare it with (36-20), we aie tempted to say that the solutions of 
(36-22), for small values ol b , can differ only slightly from solutions of 
(36-20), because the right-hand members of these equations are nearly 
equal if b is sufficiently small. The fact that this is not so is obvious from 
the foregoing remarks concerning the different behaviors of solutions of 
(36-16) for positive and negative values of b. 

Thus, in general, small changes (or perturbations) in the coefficients of 
a differential equation may completely alter the nature of its solutions. 
This remark has an important bearing on the problem of constructing 
differential equations that purport to represent the behavior of physical 
systems. In physical problems, the coefficients in a differential equation 
are usually related to physical quantities. Such quantities are determined 
from measurements which are subject to experimental errors. For this 
reason, it is exceedingly important to know just what effect small variations 
in the coefficients of a given equation have on the character of its solutions. 
When small changes in the coefficients result in small changes in the solu- 
tions, the solutions are termed stable. 



SYSTEMS OF EQUATIONS 


105 


SEC. SO] 

Another type of the stability problem occurs in the study of the depend- 
ence of solutions on small changes in the initial values. In practice one 
ordinarily seeks particular solutions that satisfy specified initial data. 
The initial data are generally determined either experimentally or from a 
specific assumption that certain physical conditions hold. (For example, 
one may assume that the deflection of a beam at a given point is zero.) 
If the initial conditions are altered slightly, is it true that the solution of 
a given equation will not be affected by a great deal? The fact that solu- 
tions of differential equations need not be continuous functions of initial 
conditions is clear from the following examples. 

Consider the solution of 

du „ 

— = ~a 2 y, a* 0, (36-23) 

at 

subject to the initial condition y( 0) ** y$. The desired solution obviously 
is 

2/(0 = y ^~ aH . 

Now, if yo is changed by a small amount A the corresponding solution is 
y(t) = (?/o + Ay 0 )e~ ah . 

Because of the factor c“~ aa< , 

1 2/(0 ~ 2/(01 0 as / co, 

and hence for any e > 0 we can choose a t 0 such that 
1 2/(0 - 2/(01 < « if t > t 0 . 

Having chosen / 0 , we let Ay be so small that 

1 27(0 - 2/(0 1 < e if 0 < / < /<>• 

Then it follows that \y — y\ < e on the whole interval 0 < / < <*>, and 
lienee the solutions are stable . By (30-5) similar arguments apply to sys- 
tems of equations with constant coefficients, and it is found that the sys- 
tem (30- 1) has stable solutious when all roots of the characteristic equa- 
tion (30-4) have negative real parts. 

On the other hand, if we solve 

dy 

— = ory, a ^ 0, 

at 

subject to the same initial condition y{ 0) = y 0 , we get 

y(t) «* yoe aH . 


(30-24) 



106 ORDINARY DIFFERENTIAL EQUATIONS [CHAP. 1 

On replacing y 0 by y 0 + Ay 0 we get 

V(t) = (2/o + Ay 0 )e a ' 1 , 

and this time \y(t) — y(t) | = e° ,( | Aj/ 0 | • This becomes infinite as t -* «o, 
no matter how small Ay 0 is, so that the solutions of (36-24) are unstable. 


PROBLEMS 


I. Use the method of this section to obtain the general solution of the system 

dy 2 

Vi 4* 2/2, -^7 - 4 y x + y 2 . 


dyi 

dt 


2 . A system of linear second-order equations 


d?yi 

dt* 


anVi + ai22/2 4 b n2/», 


d?y 2 
d/ 2 


« 2 l 2 /l 4* 0221/2 4 b 02n2/n, 


dV 

<fc 2 


* O n i2/i 4* On2l/2 4 * 4“ 0»»2/», 


where the a,j are constants, is encountered frequently in dynamics. Show by assuming 
solutions in the form 

y% ** k % cos ( \i 4- «), i * 1, 2, . . n, 
that one is led to the following characteristic equation for X: 


on 4- X 2 

Oi2 

Om 

021 

022 4- X 2 * 

02 n 

Onl 

0»2 

Ann 4- X 2 


The constants k % are determined from the system of linear equations analogous to (30-3), 
and the constant a remains arbitrary. 

8* Reduce the system of n second-order linear equations with constant coefficients, 


d?y% 

dt* 


V' , ^7 , dyj 
;-i j-i of 


to a system of 2 n first-order equations. 


t - 1, 2, » 



CHAPTER 2 
INFINITE SERIES 




The General Theory 

1. Convergence and Divergence 111 

2. Some Basic Properties of Series 116 

3. Improper Integrals and the Integral Test 118 

4. Comparison Term by Term 122 

5. Comparison of Ratios 125 

6. Absolute Convergence 127 

7. Uniform Convergence 132 

Power Series and Taylor’s Formula 

8. Properties of Power Series 138 

9. Taylor's Formula 143 

10. The Expression of Integrals as Infinite Seri as 147 

11. Approximation by Means of Taylor’s Formula 149 

Power Series and Differential Equations 

12. First-order Equations 153 

13. Second-order Equations. Legendre Functions 155 

14. Generalized Power Series. Bessels Equation 159 

Series with Complex Terms 

15. Complex Numbers 166 

16. Complex Series 169 

17. Applications 171 

Fourier Series 

18. The Euler-Fourier Formulas 175 

19. Even and Odd Functions 183 

20. Extension of the Interval 187 

21. Complex Form of Fourier Series 192 


109 





110 INFINITE SERIES ICHAP. 2 

Additional Topics in Fourier Series 

Orthogonal Functions 195 

The Mean Convergence of Fourier Series 200 

The Pointwise Convergence of Fourier Series 204 

The Integration and Differentiation of Fourier Series 207 



Although many functions encountered in applications are not elementary, 
virtually every such function may be represented as an infinite series. 

Nonelementary integrals like j (sin x 2 ) dx may be written down by inspec- 
tion as a so-called power series, and such series also give a simple, systematic 
method of solving differential equations. Another use of power series is in 
the study of functions of a complex variable z = x + iy; thus, from the 
series for sin x one can ascertain the appropriate definition and the impor- 
tant properties of sin z. A type of series known as Fourier series arises when 
one studies the response of a linear system to a periodic input, for example, 
in circuit analysis, in transmission-line problems, and in the theory of me- 
chanical systems. Fourier series and their generalizations are also useful 
for solving the boundary-value problems of mathematical physics. Inas- 
much as an indiscriminate use of series may lead to incorrect results, the 
applications presented in this chapter are accompanied by discussion of the 
circumstances in which those applications are valid. 

THE GENERAL THEORY 

1. Convergence and Divergence. A series is a sum of terms. Thus, 1 + 3 

+ 5 is a series consisting of three terms, and aj + a 2 H h is a series 

consisting of n terms. An infinite series is a series 

Ui + a 2 + fla + • • * + a n + * * * (1-1) 

which has infinitely many terms. We shall frequently use the symbol Sa n 
to denote the series (1-1). 

To get a numerical value for the expression (1-1) we consider the follow- 
ing sequence of so-called partial sums of the series, 

Si — Oi 

$2 sc ai + a 2 

«3 ** a>\ + a >2 a 3 


s n = a x + a 2 + a 3 + • • * + a n 
111 


(1-2) 



112 INFINITE SERIES [CHAP, 2 

and examine the limit of the nth partial sum s n as n — » *>, If 

lim s n » s, (1-3) 


we say that the series converges to the sum s and write 


s *= u i *T 4* ^3 + • • • + a n + * ' • • 

If the limit of s n does not exist, the series (1-1) is said to diverge , and no 
numerical value is assigned to the series. The precise meaning of the state- 
ment (1-3) is that for any preassigned positive number €, however small } one 
can find a number N such that 


\s — s n \ < € for all n > N. 
To Illustrate the definition (1-4) consider the series 

111 1 

- — -{- — - -}” — - -j- • ' * *4~ - — -f* * * 

1-2 2 3 3-4 n(n + 1) 

The first three partial sums of (1-5) are 


ai 


2 1 


, 11 2 
s 3 = s^a 3 ~l + 3 ! - - 


(1-4) 

(1-5) 


and the nth partial sum is $„ » n/(n + 1) (of. Prob 1). It is obvious that the limit 
of s n as n —► oo is 1. If, however, we want to prove this fact, we must demonstrate that 
for any preassigned number « > 0 we can find a number N such that the condition (1-4) 
is satisfied for all partial sums s n with n > N. In our problem 


\S — «n| 33 

Given « > 0, we require, then, that 

1 


1 - 


1 


and this is equivalent to 


n + 1 


n + 11 n + 1 


< f for n > N 


n 4- 1 > 


for n > N. 


Hence the choice N =* (l/«) — 1 fulfills the requirement of the definition. If e » 3d o, 
then N ■* 9; if e *= Koo» then N ** 99, and ho on. To attain higher accuracy in approxi- 
mating the sum of the seiies (1-5) by its nth partial sum s n we must, clearly, increase n. 


The number e in (1-4) can be thought of as a measure of error made in 
approximating the sum s by the sum of its first n terms. The actual error 
in the approximation is 


r« « s 


(1-6) 



THE GENERAL THEORY 


113 


SEC* 1] 

and the condition (1-4) demands that |r n | < c for all sufficiently large 
values of n. We shall call r n the remainder of the series (1-1) after n terms. 

The limit (1-3) may fail to exist either when s n increases indefinitely with 
n or when the partial sums s n oscillate without approaching a limit as 
n — > oo. Thus, the series 

1 + 1 + 1 + 14 

diverges because its nth partial sum s n — n increases with n without limit, 
while the series 

1 - 1 + 1 - 1 +••• 

diverges because its partial sums a t *= 1, s 2 * 0, s$ = 1, . . . oscillate. 

As another example, consider the so-called harmonic series 

l + H + H + /^ + H + M+ H + 3 / 8 + * — h 1/n H . (1-7) 

The terms of the series (1-7) may be grouped as follows: 

i + A + (A + H) + 04 + Ye + Yt + A) + 04 4 f He) 4 — *• 

(1-8) 

Now, each term of the foregoing series is at least as large as the correspond- 
ing term of 

Yi + A + 04 4* H) + (A + A 4~ A + Hi) + (He 4 f He) -r — . 

d-») 

The latter series, however, reduces to 

3 / 2 + /4 + K + 3^ + /4 + ’* # > 

which is divergent. Hence (1-8) is divergent. 

Tins example illustrates the idea of comparison , which is fundamental in the study of 
series. The divergence of (1-8) was established by comparing (1-8) with a simpler series, 
V T~9), whose divergence is obvious. The full chain of reasoning is as follows: “Each term 
of (1-8) is at least as great as the corresponding term of (1-9). Hence the partial sums 
of (1-8) are at least as great as the corresponding partial sums of (1-9). But the partial 
sums of (1-9) become arbitrarily large if we take enough terms. Hence the partial sums 
of (1-8) also become arbitrarily large, and the series diverges ” The student who under- 
stands this example will have no difficulty with the more detailed applications which 
follow. 

The use of the criterion (1-4) for convergence of the series (1-1) requires 
knowledge of its sum 5. F requently it is possible to infer the existence of a 
limit s without knowing its value. For example, consider the series 

0.1 + 0.01 + 0.001 + •** 

whose partial sums are s* ~ 0.1, s 2 = 0.1 + 0.01 =» 0.11, « 0.1 + 0.01 

+ 0.001 * 0.111, and so on. Each partial sum, being a decimal, is less 




114 INFINITE SERIES [CHAP. 2 

than 1. On the other hand the & n increase with n. If the successive values 
of s n are plotted as points on a straight line (Fig. 1), the points move to the 
right but never progress as far as the point 1. It is intuitively clear that 
there must be some point s, at the left of 1, which the numbers s n approach 
as limit. In this case the numerical value of the limit was not ascertained, 
but its existence has been established with the aid of a Fundamental 

Principle: If an infinite sequence of 
numbers s n satisfies the condition s n+ i 
> s n for each n, and if s n < M, where 
M is some fixed number, then s n has a 
limit that is not greater than M. In 
Fig 1 other words: Every bounded increasing 

sequence has a limit . Considering 
Sn instead of s n gives a corresponding statement for decreasing sequences. 
From the geometrical interpretation of the fundamental principle it 
appears that when an increasing sequence of partial sums s n has a limit, 
the difference between the successive values of s n must tend to zero as 
n —► oo. Since $ n — s n _ x = a n , the foregoing statement is equivalent to 
the assertion that lim a n = 0. This can be established from the defini- 

n — * * 

tion (1-3) without appeal to the fundamental principle and without the 
assumption that s n is increasing. 

Indeed, since 

ein “ $n Syi — j (1-10) 

and since the series converges by hypothesis, we have lim s n = lim $ n „i = s 
as n — > oo. Hence (1-10) shows that 

lim a n = lim s n — lim s n ^i = 0. (1-11) 


We state the result (1-11) as a theorem: 

Theorem I. If a senes converges , then the general term must approach 
zero , and hence if the general term does not approach zero, the series diverges. 

The reader is cautioned that the converse of this theorem is not true. 
For instance, the harmonic series (1-7) was found to diverge even though 
the general term 1/n approaches zero. 

There is a more elaborate version of Theorem I which does have a converse. By 
writing out the sums m full we find a relation analogous to (1-10): 

Om + 1 H h On s «« - \ t 71 > m > 1 . ( 1 - 12 ) 

If the infinite scries converges, so that lim s n * s, then both the sums on the right of 
(1-12) become arbitrarily close to s, provided ?n and n arc chosen large enough. Hence 
the right-hand side becomes arbitrarily small in magnitude, and we are led to the follow- 
ing: IfXak converges , then for any e > 0 there is an N such that 

\Om 4 a m + 1 4 * b ttnl < « 


(1-13) 



THE GENERAL THEORY 


115 


SEC. 1] 

whenever n > m > N. Now this statement admits a converse. 1 //, for each e > 0, 
there is an N such that (1-13) holds whenever n > m > N, then Xak converges. The 
theorem, together with its converse, constitutes the so-called Cauchy convergence criterion . 

Example 1. A certain series has partial sums s n » r n , where r is a constant such that 
0 < r < 1. By use of the fundamental principle, show that the series converges to zero. 
We have to show that lim «„ *■ 0 as n — ► or in other words 


limr n 


0 


for 0 < r < 1. 


(1-14) 


Since r > 0, it is evident that s n > 0, and hence the sequence s n is hounded from below. 
Also r n+1 « rr n , or in other words 

rSn- ( 1 - 15 ) 

Since r < 1, this shows that s n -i~i < s«, so that the sequence *« is decreasing. Hence the 
limit of 8 n exists by the fundamental principle. If we write s » lim s n and take the 
limit as n — ► « in ( 1 - 15 ), there results 

8 « lim a„+i « lim ( rs n ) « r lim s„ *■ rs. 

From s ■» rs it follows that s 0, since r ^ 1, and this gives (1-14). 

Example 2. The geometric series is defined by 

1 + x 4 x 2 + x 3 H + x n H . 

Show that this series converges to 1/(1 — x) when |x| < 1 but diverges 
when | x j > 1. 

The geometric series is an example of a series 

u x {x) 4 u 2 (x) + u 3 (x) H b u n (x ) 4 • • * 

in which the terms are functions of x. For each choice of x the function 
u n {x) is simply a number, the series becomes a series of constants, and 
hence it can be tested for convergence just as any other series of constants 
is tested. 

We have to decide whether the partial sums 


s ft =l + x + z 2 +"4 * n ~ 2 4 x”- 1 (1-16) 

tend to a limit. If the foregoing equation is multiplied by x, there results 


xs n 


x 4 x 2 H b x n 4 x n 


(1-17) 


and subtracting (1-17) from (1-16) yields s n — xs n = 1 — x n . Solving for 
s n we get 

1 — x n 

Sn m (148) 


1 — x 

1 Since we shall not require the converse, the proof is not presented here. The inter- 
ested reader is referred to I. S. Sokolnikoff, “Advanced Calculus,” pp. 11-13, McGraw- 

Hill Book Company. Inc., New York, 1939. 



[chap. 2 


116 INFINITE SERIES 

[f |$| < 1, then lim \x\ n « 0 by (1-14) and hence (1-18) gives 

» -+ ao 

i-o 1 

iim s n = ** * 

n « 1 — x 1 — x 

This establishes the required convergence when \x\ < 1. On the other 
hand if |»| > 1, the general term does not approach zero and the series 
diverges by Theorem I. The value x is called the ratio for the series, since x 
equals the ratio of two successive ter mk. We have shown that the geometric 
series converges if, and only if, the ratio is less than 1 in magnitude. 

PROBLEMS 

1. Show that the nth partial sum of the series (1-5) is n/(n -f 1) Hint' Since 
l/[w(T4 4* 1)] * l/n — l/(n -f 1), the sum of the first n terms is 

Sn - (H -H) + (A - H) + (U -H)+--- + [1/W - l l(n 4* 1)1 

A series such as this is called a telescoping series . 

2. Show that the following series converge to zero if |r| < 1 but to 1 if r « 1 Sketch 
the graph of the sum as a function of r: 

r 4- (r 2 - r) 4- (r z - r 2 ) 4* • • 4- (r n - r n ~ 1 ) + 

2. Some Basic Properties of Series. We .shall write infinite series in the 
condensed notation 

oo 

a n 555 #1 4“ "b a 3 + * ’ • + U M 4" * * * . (2-1 ) 

n* 1 

Finite sums are expressed similarly, with the limits of summation (1 ,qo) 
replaced by the appropriate values. The limit s of summation are frequent ly 
omitted if they need not be emphasized or are clear from the context. 
Whenever the limits are omitted in Sees. 2 to 7 of this chapter, the reader 
may assume that the summation range is from 1 to oo. 

In many respects convergent series behave like finite sums. For example, 
if the sum of the series (2-1) is vS and if each term of the series (2-1) is multi- 
plied by a constant p, then 

2pa n = p2a n ~ ps. (2-2) 

That is, a convergent series may be multi plied termwise by any constant. 

The proof of (2-2) follows at once from die observation that the partial 
sums & n of the series 2pa n are related to the partial sums s n of (2-1) by 

S n - P*n 

lim S n * p lim s n = ps. 


and therefore 



THE GENERAL THEORY 


117 


8EC. 2 ] 

If we are given two convergent series 2a* and 26*, then 

2 (a* ± 6*) — 2a* dh 26*. (2-3) 

That is, two convergent series may be added or subtracted term by term. Again 
the proof is simple. We denote the sum of the series 2a n by A, that of 26 n 
by B, and the corresponding partial sums by A n and B n . Then the nth 
partial sum of 2 (a* zb 6*) is 

n 

4: ” A n zb B n 

k*~l 

and the result (2-3) follows on letting n —► qo. 


As an illustration, consider the geometric series 


By (2-2) 


:rs » x 4* x 1 -f • • • -f r n 4 


and hence, by (2-3), we have * — jts =» 1. This shows that if the series converges , it must 
converge to 1/(1 — x). The question of convergence was discussed m Example 2 of 


Sec. 1. 


Another obvious but important property is used so often that we state 
it as a theorem: 

Theorem I. If finitely many terms of an infinite series are altered the 
convergence 'is not affected ( though , of course, the value of the sum may be 
affected). 

To prove this we denote the original terms by a* and the new terms by 
a* + 6*, where all but a finite number 1 of 6*s are zero. The result is then 
a consequence of (2-3). It should be noticed that this argument not only 
establishes convergence but shows that the new value of the sum can be 
found by the obvious arithmetical calculation. For instance, if the seventh 
term of a convergent series is increased b}" 2.1 the sum is also increased by 
2.4, and similarly in other cases. 

Example: Establish the divergence of 

h.2 + he 4- ho 4* Via 4 • (2-4) 

Multiplying by 4 we get the senes 

Vs + X + H + H 4 — * 

which is obviously divergent, since it differs from the harmonic series 2l /n only in that 
it lacks the first two tcirtis. Hence (2-4) is divergent. 

1 Any finite series h 4- ■ * • 4* &« may be regarded as an infinite series with all terms 
beyond the nth equal to zero. If wo do so regard it, the definition of convergence given 
in Sec. 1 makes the finite series converge to its ordinary sum. 



118 INFINITE SERIES [CHAP. 2 

This use of (2-2) to establish divergence is readily justified, even though (2-2) applies 
to convergent series only. Thus, assume that (2-4) converges. The foregoing analysis 
shows, then, that Si fn would have to converge, and that is a contradiction. 


PROBLEMS 

1 . Write the following series in full, without using 2 notation: 


^ 1 ~ / 43\» n 2 + 1 ^ 3 

hi 2k' h\ U/'hi'Zn + Z hj' hi 


(cos x) . 


2 . Write the following series in condensed form, using 2 notation: 


(H) 2 4-(H) 3 + (H) 4 4-(H) 5 4-**-, 
T[o55 + 1^002 + 1^004 + 1^006 + ’ ’ 

1 + -L + -L- 

11-2 1-2-3 1-2-3-4 


H + H 4- M 2 4 - Ms 4-* • 


0.2 - 0.02 + 0.002 


Ho + Ho + Mo + Mo 4 • 


3. Some of the series in Probs. 1 and 2 are divergent because the general term does 
not approach zero. Which ones are they? 

4 . Some of the series in Probs. 1 and 2 are convergent because they are geometric 
series with ratio less than 1 in magnitude (or multiples of such a series). Which ones 
are they? 

6. Some of the series in Probs. 1 and 2 are divergent because 2l/n is divergent 
Which ones are they? 

0. Shovr that (1 — 1) 4- 0 — 1) 4- (1 — 1) 4 converges but would diverge if the 

parentheses were dropped. 

7. (a) Does the series 2 bb+©] converge? Explain, (b) Does the 

converge? Explain. Hint In (a) see (1-5). In (6) note that 
HMD] 4* . If the given series converges, what could you deduce about 


3. Improper Integrals and the Integral Test. In the development of the 

fb 

calculus a definite integral such as / f(x) dx is defined, at first, only for a 

Ja 

finite interval [a,b]. The extension to an infinite interval is then made by 
a simple passage to the limit; thus 


f f(x) dx — lim [ f(x) dx . 

Ja h — * ao a 


The integral at the left of (3-1) is called an improper integral. If the limit 
at the right exists, we say that the improper integral converges (to the value 



SBC. 3J THE GENERAL THEORY 119 

of the limit) and it diverges if the limit does not exist. The definition is 
quite analogous to the corresponding definition 

*> n 

2^= lim J2 a k 

k«*l n O0 kw~l 

for infinite series. 

An example of a divergent improper integral is 

r -dx ** lim / ~ lim (log x |}) «« lim log b «•* (3-2) 

J i x J i x 

On the other hand if p is constant and p s* 1, then 

r \dx - lim dx - lim f) - lim — - — - ■ (3-3) 

/i x p J i VI — p If/ 1 — p 

The question of convergence now depends on the behavior of b l ~ p as b — ► «>. If the 
exponent 1 — p is positive, then b l ~ p — ► « and the integral (3-3), like (3-2), is divergent. 
But if 1 — p is negative, then p — 1 > 0 and hence 

b l ~ p **= —► 0, as b — * oo. 

In this case the integral (3-3) converges to the value l/(p — 1). 

The result of this discussion may be summarized as follows: 

r* 1 

Theorem I. The improper integral / — dx converges if , and only if, the 

Ji x v 

constant p > 1. 

Theorem I suggests the following analogous result for infinite series: 

00 i 

Theorem II. The infinite series ^2 ~~~ converges if , and only if, the con- 

k-i k p 

slant p > 1 . 

It will be seen that Theorem II is valid ; in fact, there is a close connec- 
tion between infinite series and improper integrals which will now be dis- 
cussed. 

Suppose the terms of an infinite series 2 a* are positive and decreasing; 
that is, a n > a n+ i > 0 for each positive integer n. In this case there is a 
continuous decreasing function f(x) such that 1 

&n 355 /(r)j n — 1, 2, 3, .... (3-4) 

Each term a n of the series may be thought of as representing the area of a 
rectangle of base unity and height f(n) (see Fig. 2). The sum of the areas 

1 For instance, let the graph of y ** f{x) consist of straight-line segments joining the 
points (n,o») and (n -f- 1, a n +i)- 



INFINITE SERIES 


120 


[chap. 2 


of the first n circumscribed rectangles is greater than the area under the 
curve from 1 to n + 1, so that 


rn+l 

a t + a 2 -\ h a n > J f(x) dx. (3-5) 

/ 00 

fix) dx diverges , then the sum 2 Jo* also 

diverges. 

On the other hand, the sum of the areas of the inscribed rectangles is 


iv 



Fiu. 2 


less than the area under the curve, so 
that 

02 + 03 + • * * + ^ f x fi x ) dx. ( 3 - 6 ) 

If the integral converges, we have [since 
fix) > 0 ] 

^ fix) dx < f(x) dx ss M, 

so that the partial .sums are bounded 
independently of n: 

$„ = of! + a 2 H a n < M + a x . 


Since each a* is positive, these partial sums form an increasing sequence. 
Hence, the fundamental principle stated in Sec. 1 ensures that ~a k is con- 
vergent. 

The result of this discussion may be summarized as follows: 

Theorem III. For x > 1 let fix) be positive, continuous , and decreasing. 

00 r° 

Then the series Y) fin) and the integral / fix) dx both converge or both di- 
rt- 1 h ' 

/n either case the partial sums are bounded as follows: 

n + l n n 

f SO) rfx < E SO) < fjO) dx + SO). (3-7) 

Jl lea*. 1 

Choosing /(x) = x~“ p in Theorem III, we see that Theorem II is a conse- 
quence of Theorem 1. The test for convergence contained in Theorem III 
is commonly called the Cauchy integral test , though it was first discovered 
by Maclaurin. The result (3-7) is especially useful because it enables us 
to estimate the value of the sum. 

Example 1 . Show that the series 

_J_ + _J__ + __L_ + _JL_ + . . . + _L_ + . . . 

1 + l 2 ^ 1 + 2 2 1 + 3 5 1 + 4 2 ^ l+n* 

converges to a v^lue which is between 0.7 and 1.3. 


121 


SEC. 3] THE GENERAL THEORY 

Here we choose /(x) ■» 1/(1 -f ar 1 ). Since 

1 , . . ,fc r ** r 

j as 6 -* *>, 

the integral is convergent, and hence the series is convergent. Moreover 

„ _ IT * 1 7T 1 

0.79 * 7 < E 7-T-7;, < 7 + - * 1.29 
4 m 1 -h k 2 4 2 


C 


by letting n — ► « in (3-7) and noting that /(l) * V 2 . The next example shows how the 
accuracy in such an estimate may lx? improved to any extent desired. 

Example 2. Compute the sum of the following series within JrO.Ol: 


1 •— d — ~f -f- - ~ -j- — -4" • 

4 9 lti 25 36 


n 1 


It is easily verified that the first six terms give the sum 1.491. To estimate the re- 
mainder we have, from (3-7) on taking /(x) *= 1 /(x 4 6) 2 , 

f (xTo ? dl < ? («+>? < .(V + 0 ? dl + S' (3_8) 

The two limits in (3-8) are 0 143 and 0 163, as the reader can verify. Hence 

1.634 - 1.491 + 0.143 < « < 1 491 + 0 163 « 1.654. (3-9) 

It is interesting to see how many terms are needed to get the same accuracy by direct 
computation. The remainder aftei n terms is given by (3-7) as 

t f \<ix- -• 

Jn+iX* n + 1 

To make tins as small as the uneeitainty interval 1 054 — 1 634 obtained in (3-9), we 
must have l/(n -f 1) < 0 02, or n > 19 Thus, direct summation of the series requires 
almost 50 terms for the aocuiaey which we obtained by adding 6 terms only. 


PROBLEMS 

1. Test the following integrals for convergence, and evaluate if convergent: 

dx dx 


r r e -, dXt /v.*, r_* r 

Ji 1 + x Ji Ji h x(log x) 2 J 2 . 


x logs 


2 . Test the following series for convergence: 

1 ^ A 1 


' <n + 1) H 


1 y' 1 y I v 

’ n»2 n(Iog n) ^2 n(log n)' 01 w 1 -f n 2 


3 . (a) For what values of the constant c does j^e CT dx converge? ( b ) Using the 

result (a), discuss the convergence of 2c cn . (r) Show that the series (h) is a geometric 
series, and also show that your results are consistent with those of See. 1. 

4. How many terms of the harmonic series Sn - " 1 are needed to make the sum of those 
terms larger than 1,000? 



INFINITE SERIES 


122 INFINITE SERIES [CHAP. 2 

00 

5. Estimate the value of £ n~ 4 by direct use of Theorem III and also by adding the 

»«*i 

first five term® and using Theorem III to estimate the remainder. In both cases find 
approximately how many term® of the series you would have to add up to get comparable 
accuracy. 

Problem for Review 

6. (a) By (1-18), show that the partial sums of the series 

1 + 2 + 2* + 2*~l •" 2 " + "' 

are all less than 2. 

(6) Show that the partial sums of the series 

i -fJL-i._L-t-.I-i — lIj — 

21^3! 4!^ n!^ 

are also less than 2. Hint * Compare the partial sums with those of the series (a), 

(c) Deduce, by the fundamental principle, that the series (6) converges. 

4. Comparison Term by Term. One way to test a series of positive 
terms for convergence is to compare that series with another whose con- 
vergence is known. Let 2a n and 2b n be two series with positive terms such 
that a n < b n and 2b* converges. The inequality 

n n ac 

®n « £ <>n < £ b n < £ b n 

1 1 1 

shows that the partial sums $* are bounded, and since s n is increasing, the 
limit exists by the fundamental principle. It is left for the student to verify 
also that if a* > b n > 0 and 2b* diverges, then 2a n diverges 

This discussion establishes the following result, known as the comparison 
test: 

Theorem I. If 0 < a n < b*, then the convergence of 2a n follows from the 
convergence of 2 b n . And if a n > b n > 0, then the divergence of 2a* follows 
from the divergence of 2b„. 

Since the first few terms of a series do not affect the convergence, we need the hy- 
pothesis not for all n but only for n sufficiently large (see Sec. 2, Theorem I). Similar 
remarks apply to every convergence test, and we shall make constant use of this fact in 
the sequel. 

For example, suppose we want to establish the convergence of 29/n n . Although the 
inequality 

9 1 


is not valid for all n, it is valid when n is sufficiently large. Hence the series converges by 
comparison with the geometric series. Another example is given by the series 

„?2 100 log n' (4 ~ 2) 



BBC. 4] 

Although it is not true that 


THIS GENERAL THEORY 


123 


1 1 

100 log n n 

for all n, this is true for all sufficiently large n, and hence the series (4-2) diverges by com- 
parison with the harmonic series. 

It is customary to write a„ ~ b n (read “a n is asymptotic to b n ”) if 


<Z n 

lim — — 1 

n -+ *>b n 

(compare Chap. 1, Sec. 2). For example, n + 1 ^ n and also 5 n 2 + 3n 
+ 4 ^ 5n 2 , but it is not the case that 2/n ~ 1/n even though the difference 
between these quantities tends to zero. In this notation we can state the 
following theorem, which is very useful for determining convergence: 

Theorem II. If a n ~ b n and b n > 0, then the series Xa n and Xb n are both 
convergent or both divergent. 

The proof is simple. Since lim ( a n /b n ) = 1, we shall have 


1 a n 

- < — < 2 

2 b n 


whenever n is sufficiently large. Equation (4-3) yields 

Vzb n < a n < 2 b n 


(«) 


and hence the conclusion follows from Theorem I together with (2-2). 

Example 1. DoesSw -10 *" converge? 

For all large n we have log w > 2 (since log » — * «). Hence 


for all large n, and the series converges by comparison with the convergent series Xl /n 2 
(Theorem II, Sec. 3). 

Example 2. Does X(n 2 -f 5n + 3)“** converge? 

Inasmuch as n 2 -f bn -f 3 *« n 2 (l -f 5/n + 3/n 2 ) ^ n 2 , we liave 

(n 2 + 5n + 3)“* ~ (n 2 )~* - n~‘ - ~ 

n 

Since Si /n diverges, the given series diverges. 

Example 3. Consider the series 

_ / ra 4 + 4n» + 1 
\7 n 1 + 5n* + 8 n) 


Since n* + 4n* + 1 ~ n 4 , and since 7n 7 + 5n 4 + 8n 

/ n 4 \H J_ _1_ 
\7nV ”7 m t»W 




totic to 


i n , tue general 




— — r 


The series with general term 1 /n* converges by Theorem II, Sec. 3, and hence the given 
series also converges. 



124 


INFINITE SERIES 


[CHAP. 2 


Examples 2 and 3 illustrate two properties of the relation which 
are now set forth explicitly. First, we show that any polynomial is asymp- 
totic to its leading term. Indeed, if a 0 and m > 1 , then as n °o, 


an m + bn m 1 H — • + rn + s b 

z * 1 + " — I f~ 


an 


an 


an 


tn— 1 


+ 


— > 1. 


an 


This shows that an m + bn m 1 H 1- rn + s ^ an m , as stated. 

Second, if a n ^ b n and c n ~ d n , then it follows that 

~ bldi 


for any constants a and To establish this consider the ratio 


= /On\ a /cA* 

b a X \bj \dj 


in' 3 = l. 


PROBLEMS 

1. Test the following series for convergence by comparing with the series 2l/a p : 

v _£ y 1 V ,lS v _Jl 

V r t 2ny/ n + 1 (2« + l) J (2» -I- 1)* 

2. Test the following series for convergence by using Theorem It : 

s 5l± A vAt "!, v (l , J y v 3 1+J!, v «l±I. 

n 3 5 + 1 3n 6 -H n \n n 2 ) 4 n -f 5 n a 4 -f 4 

3. Test the following series for convergence by any method: 

1 In 4 

y.-n* V f V . } V 

’ w log (n -f 1) ~ n^” w « 6 -j- 3 

4. (a) If a n ~ ?> n and b n ~ c n , show that a n ~ c* (/>) If a n ~ h* and c n ~ d n , is it 
necessary that a„ t r» ^ h n -f ri n ? Prove your answer by an eximple. (r) Find a n 
and 6 n such that a n ~ b n but a n — b n — > oo {<!) Find a n and b n such that a n /b n <*> 
but a n — b n 0. 

Problem for Review 

5. (a) By direct use of the definition of limit show that 0.111111 ...= Hint * 

If *» 0.1, 82 33 0.11, ss * 0.111, ...» then |si — %\ « l«2 — I 33 i6oo» and 

SO on. (6) With s n as in (a), and with e > 0, how large must you choose N to make 

|s« — H I < « for all n > N? 

(c) If s **0.111111..., evaluate s by considering 10s — s. (d) Evaluate s in (c) by the 
formula for sum of a geometric series. 



THE GENERAL THEORY 


125 


SEC, 5] 


5. Comparison of Ratios. It often happens that the general term of 
an infinite series is complicated whereas the ratio of two successive terms is 
simple. For example, in the series 


we have 



a n 


x 2n + 2 n\ _ x 2 
(n + 1 ) ! x 2n n + 1 


(5-1) 

(5-2) 


The following theorem enables us to deduce convergence by considering 
this ratio rather than the general term itself: 

Theorem I. Let 2a n and be two series with positive terms . If 


a n + 1 bn 41 

cifi bfi 


1 , 2 , 3 , 


(5-3) 


then the convergence of Za n follows from the convergence of And if 


Gn-f 1 

a n 



* = 1.2,3, 


(5-3a) 


then the divergence of 2a n follows from the dierrytnee of Zb n . 
The proof is simple. Tn the first case we have 


a 2 a 3 a n b 2 b 3 

a n = ai < <h ~ - 

a i « 2 <*n~i fh b 2 


b n a x 

SB — l ) 

bn - 1 b x 


Hence the convergence of 2S5 n implies that ol 2 a u by the comparison test 
(Theorem I, Sec 4). The discussion of (5-3a) is similar. 

If we take b n — r n in Theorem I, then ~b n converges whenever r < 1. 
Also 


bn + 1 


j.n ~H 


r. 


b 


n 


r 


n 


Hence the theorem shows that 2a n converges if there is a fixed number 
r < 1 such that 

i»-l,2,3, .... (5-4) 

Cl n 


Since the condition 
verges whenever 


(5-4) is needed only for large n, the series 2a n also con- 


lim =* r < L 

n « a n 


(5-5) 



126 INFINITE SERIES [CHAP. 2 

The test based on (5*4) and (5-5) is termed the ratio test. To illustrate 
the ratio test consider the series (5-1). By (5-2) we have 

,. Gn+l 

lim = lim = 0 

a n n + 1 

and hence (5-5) holds for all x. Thus the series (5-1) converges for all x. 


The ratio test is useful but very crude. It cannot even establish the convergence of a 
series such as 2n~” 1(K) , which is rupidly convergent. To obtain a better test one may use 
the series 2l/n p for 26 n rather than the geometric series In this case 


frn+l 

&n 


1 

vn 4- 1)* 


n v 


c-tt)*- o+r- 


By the binomial theorem 1 


and hence 



V , P (P 4- 1) 
n 2 n 2 


bn±l _ ^ _ P 

b n n 


Since Xb n converges if p > 1 and diverges if p < 1, we are led to the result stated in 
part (6) of Theorem II. The result (5-5) is stated in part (a) 


Theorem II. Let Xa n be a series of positive terms, and let r arui p be con- 
stant . (a) If a n + 1 /a„ ^ r, then Xa n converges when r < 1 and diverges when 
r > 1. (b) If a n +i/a u — 1 ~ -~p/ri , then 2a n converges when p > 1 and 

diverges when p < 1. 


Example 1. Does 2n 2 /2 n converge? 
With On 30 n 2 /2 n we have 

fln+i (a 4" 0~ 2” 
~aZ ** ”7? 






Hence the series converges by the ratio test, Theorem I la. 
Example 2. Apply the ratio test to the harmonic series. 
With On * 1/nwe have 

fln-fl ^ n ^ ^ 

a„ n + 1 


Since this is the case r « 1, the test gives no information. Moreover, 


fln-f l ] cs 71 

a n n 4 1 


1 1 
-- - » • — — * 

n 4 1 n 


Since this is the case p «- 1, the more refined test of Theorem 116 also gives no infor- 

mation.* 


1 The binomial theorem for arbitrary exponents is established in Sec. 12. 

s More general tests may be found in I. S. Sokolnikoff, “Advanced Calculus/' chap. 7, 
McGraw-Hill Book Company, Inc., New York, 1939. 



SEC. 6J THE GENERAL THEORY 127 

Example 3. For what values of the constant c does the following converge? 

c . c(c + 1) . c(c + l)(e + 2) , 

— •+•« 

1! 21 3! 

For sufficiently large n the terras are of constant sign, and hence Theorem II is ap- 
plicable. We have 

firm c(c 4- l)(c 4- 2) ... (c 4- w) n* c + n 

a n (n -f 1)! r(f 4- 1) . . . (c + n - 1) n + 1 

Since 

c -f n ^ ^ ^ c - 1 c ~ 1 ^ ^ 1 - c 

n + 1 n + 1 n n 

the series is convergent if l — c > 1 and divergent if 1 ~ c < 1. Hence, it is convergent 
when c < 0 and divergent when c > 0. In this example Theorem I la gives no informa- 
tion but Theorem 116 solves the problem completely 


PROBLEMS 


1. Determine the convergence by using the ratio test, Theorem Ila: 


X 



n! 

n* 



v fr 2 -f l) n 

W ’ n! 


2. Show that Theorem Ila gives no information, and test for convergence by The 
orem 1 16; 

v } ? v *}' f v l 2 alL. 

n 2 ** r(c -j~ l)(c -f- 2) . . . (c + n) 4 n (n!) 2 

3. Test for convergence by any method : 

• n<_ log n, v 2"_+JP f v 2n - 1 

1*3*5... (2n -4-1) n 2 * w 3 n + 4"' " 2n +T 

4. Give an example of a divergent series Xan such that all the terms satisfy a n > 0 
and a n+ i/a„ < 1. Does this contradict the remarks made m connection with (5-4)? 

5. If j is constant prove that lim x n /n\ » 0. 11 ml The series X j x\ n /n' converges 

n — * » 

by the ratio test. 

6. Absolute Convergence. The preceding tests for convergence apply 
to series with positive terms. We shall now see how these tests can be used 
to establish convergence even when the signs of the terms change infinitely 
often. 1 

Definition. A series Xa n is said to he absolutely convergent if the series 
of absolute values X | a n | is convergent . 

1 If all but a finite number of terms have the same sign, then we mMy consider those 
terms only (Sec. 2, Theorem I). Multiplication by — 1, if necessary, yields a series with 
positive terms, so that the foregoing methods apply. This fact was used in Example 3 
of the preceding section. 



128 


INFINITE SERIES 


[CHAP. 2 


For example, the series 2 (~~l) n /n 2 is absolutely convergent, since 


ZIa*| 



2 


converges. On the other hand the series 2( — l ) n /n is not absolutely con- 
vergent, as the reader can verify. The importance of absolute convergence 
sterns partly from the following theorem: 

Theorem I. 7/2|a n | converges, then 2a u converges . 

In other words, an absolutely convergent series is convergent . The defini- 
tion of absolute value yields 


0 < 


a n + 

2 



Hence, by the comparison test, the series 

v (l n + i «n | 
Zj 

2 


converges when 2la n | converges And then the series with general term 


a n — 



converges by (2-2) and (2-3). 

To illustrate the use of Theorem I consider the series 

v cos nx 


( 6 - 1 ) 


Since the signs change infinitely often, 1 none of the preceding methods is applicable 
We may, however, apply those methods to the series of absolute values. In view of the 
fact that 

| ros nx | 1 

— - ~ C ~ ? 

2« — 2 n 


the series 

is convergent. Hence the original series ((>-1) is convergent. 

A series whose terms arc alternately positive and negative is called an 
alternating series . There is a simple test due to Leibniz that establishes 
the convergence of many such series even when the series does not converge 
absolutely. 

1 Except when x is an integral multiple of 2 r. 


ros nx 


cos nx 
2" 



129 


BSC. 6] THE GENERAL THEORY 

00 

Theorem II. Suppose the alternating series 53 (-l) n+1 a„ is such that 

n—l 

a n > a n+1 > 0 and lim a n = 0. Then the, series converges , and the remainder 
after n terms has a value which is between zero and the first term not taken. 

For example, if the sum of the series is approximated by the first five 
terms 

8 * $5 = eii — a 2 + — a± + 

then the error in that approximation is between zero and -~oq : 

0 > 8 — $5 > — Ofi. 

The value given by is too large, because s r> ends with a positive term, +a 6 . 
The value sq is too small, since ends with a negative term, and so on. 

To prove the theorem, we have 

* <ai — O2) + (a 3 — 04) H b (o^n-l — 02 n) 

“ <*1 («2 “ «3) ~ * ~ (02n-2 ~~ «2n-l) “ «2n 

and hence $2„ is positive but less than a 1 for all n. Also 

82 < 84 < 8 b * . . 

so that these sums tend to a limit by the fundamental principle (Sec. 1). Since ** 
«2n -f a2 n -f i and lim a2M \ 38 0, it follows that the partial sums of odd order tend to 
this same limit, and hence the series converges. The proof of the second statement is 
left as an exercise for the reader. Actually Theorem II becomes rather obvious when we 
plot the partial sums on the t axis. 


Since the choice a n — 1/n satisfies the requirements of Theorem II, the 
alternating harmonic series 


11111 (~l) n+1 

*« 1 ~;; + o - - + + • 

2 3 4 5 6 n 


(6-2) 


is convergent. If the sum is approximated by the first two terms, then 
Theorem II says that the error is between 0 and that is, 0 < s — 3^ 

< H, or 

< 8 < %. ( 6 - 3 ) 


Inasmuch as the series of absolute values diverges, we could not establish 
the convergence by use of Theorem I. A series such as this, which con- 
verges but not absolutely, is said to be conditionally convergent 

By rearranging the order of terms in a conditionally convergent series, one can make 
the resulting series converge to any desired value. In illustration of this fact we shall 
rearrange the series (6-2) in such a way that the new sum is w, though (6-3) shows that 
the original sum is not 



[chap. 2 


130 INFINITE SERIES 

The terms of (6-2) are obtained by choosing alternately from the series 

l + X + H + K +■■■ (6-4) 

and from the series 

-H - H - H - 14 («) 

both of which are divergent. To form a series that converges to it, first pick out, in order, 
as many positive terms (6-4) as are needed to make the sum just greater than 7 r. Then 
pick out, in order, enough negative terms (6-5) so that the sum of all terms so far chosen 
will be just less than ir. Then choose more positive terms until the total sum is just 
greater than ir, and so on. The process is possible because the series (6-4) and (6-5) are di- 
vergent; the resulting series converges to r because the error is less than the last term 
taken. 

To get a physical interpretation of this result, suppose we place unit positive charges 
P at the points 

x - i, -V2, Vz, -VI, VS, -VS 

and attempt to find the force on a unit, negative charge N located at the origin (see Fig. 
3). By Coulomb's law two opposite unit charges a distance \/n apart experience an 


— V8 — V© — V4 —V2 0 fl V3 fS V7 

Fig. 3 


attraction of magnitude 1/n. Since the attraction of charges at the left of N exerts a 
force toward the left whereas attraction of the other charges exerts a force toward the 
right, the total force on N is given formally by the series (6-2). Now, the fact that 
this series is conditionally convergent makes the force dejiend not only on the final con- 
figuration of charges but also on the manner in which the charges were introduced. If 
we obtained the final configuration by putting 10 charges at the left, then 1 at the right, 
then 100 more at the left, and 1 again at the right, and so on, the net force will be di- 
rected toward the left. But if we had a preponderance of charges at the right while 
setting up the final configuration, then the final force would be directed toward the right. 

The foregoing behavior is perhaps not very surprising. What is surprising is that a 
rearrangement such as this will always give the same value provided the series m ques- 
tion is absolutely convergent. For example, let the configuration consist of unit positive 
charges P at the points x * 1, —2, 3, —4, 5, —6, . . ., so that the force is given by the 
absolutely convergent series 


1 1 1 

1 -i5 + p-?+'-- + 


(- 1 )” 


In this case, as we shall show, the force does not depend on the way in which the final 
configuration was reached. 1 

The preceding examples may assist the reader to appreciate the following 
theorem, which describes what is perhaps the most important property of 
absolute convergence. 

Theorem III. The terms of an absolutely convergent series may be re- 
arranged in any manner without altering the value of the mm. 

1 One may say that the “charges at infinity" now have no influence, whereas in the 
former case (6-2) they were important. 




THE GENERAL THEORY 


SEC. 6] 


131 


We establish this result first for series of positive terms. Let 2p* be 
such a series and Xp k a rearrangement. For every n we have 


fc-1 Jfc-1 

inasmuch as each term p k is to be found among the terms of 2p*. Hence 
2 p k converges (by the fundamental principle), and also 

Spi < 2 p k . 

In just the same way we find 2 p k < 2pi, and hence 2p* = 2p*. 

To obtain the result for an arbitrary but absolutely convergent series 
2a* f denote the rearrangement by 2a* and observe that 


a k » (a k + \a k \) - |a*| 
a k » (a k + |ai|) - |a*|. 


(6-6) 


By the result for positive series we have 

2|<4l - S|a*| 

2 (a k + | ail) == 2(a* + |a*|). 

Hence (6-6) gives 2a* = 2a* when we recall (2-3). 

By methods quite similar to the foregoing 1 one can establish the follow- 
ing, which expresses a third fundamental property of absolutely convergent 
series: 

Theorem IV. If 2 a k — a and 26* ~ b are absolutely convergent , then 
these series can be multiplied like finite sums and the product series unU con- 
verge to ab. Moreover , the product series is absolutely convergent , hence may 
be rearranged in any manner . For example, 


ab = aibi + (a$bi + a^) + (a^bi + 0262 + 0163) + • • *. 


Example: Consider the series 2j n / \f 1 . 
With a* x n /y/n we have 

|gn-fil _ 1 x n + l Vn | 
I On I I Vn + 1 X n I 


1*1 




Hence the series converges absolutely if |ac| <1 and diverges if |x| > 1. To see what 
happens when x ■* dbl, we substitute these values into the original series, obtaining 


2 


(~l) w 

Vn 



for x « —1 and x «* +1, respectively. The first series is conditionally convergent, 
and the seoond is divergent. Hence the series converges absolutely when [x| < 1, it 
converges conditionally when x — — 1, and it diverges for all other values of z. 

1 The proof is given in full in Sokolnikoff, op. cit., pp. 242-244. 



VS2 


INFINITE SERIES 


(CHAP. 2 


PROBLEMS 


1. Classify the following series as absolutely convergent, conditionally convergent, or 
divergent: 


‘(HIT V, lv , g"+ 8 V HL‘ v w 1 

' Vn’ 2 n" ' " 7? ’ w 2 V 3 


3*6 3-6-9 


1 - 3 • 5 • 7 


+ • 


3*6*9*12 

2. Determine the values of x for which the following series are absolutely convergent, 
conditionally convergent, or divergent: 

S(-l)" S(-l) n ff y 2(-l )V, 2 L, S - (-i. r --)"2n!x" ( 2 

n (2n) f nx n n \i + 4/ log (n -f 1) 

eo 

3. Approximately how many terms of the series JZ ( — l)”/?i 4 are needed to give the 

i 

sum within IQ" 8 ? Evaluate the sum to two places of decimal* 


7. Uniform Convergence. If a finite number of functions that are all 
continuous in an interval 1 [a y b] arc added together, the sum is also a con- 
tinuous function in [a y b\. The question arises as to whether or not this 
property will be retained in the case of an infinite series of continuous 
functions. Moreover, it is frequently desirable to obtain the derivative 
(or integral) of a function fix) by means of term-by-term differentiation 
(or integration) of an infinite series that defines /( j). I’nfortunately such 
operations are not always valid, and many important investigations have 
led to erroneous results solely because of the improper handling of infinite 
series. The analysis of these questions is based on a property known as 
uniform convergence , which is now to be described. 

CO 

If a series of functions u n(?) converges for each value of x in an inter- 

n S3* 1 

val [a, 6], then the sum defines a function of x, 

s(x) *= Si4 n (x). 

We denote the nth partial sum by *s„(x), 


*»(*) = Ui(x) + U 2 (x) + V H (x) -f 1- Unix), 

and the remainder after n terms by r n (x): 

r n (x) = s(x) - 8„(x) =* Mn+lW + U ni *{*) H * (7-1) 

Since the series converges to s(x), lim s n (x) ~ s(x) as n oo, and hence 

lim r n (x) = 0. (7-2) 

The statement embodied in (7-2) means that for any preassigned positive 
number e, however small, one can find a number N such that 

|r n (x)| < e for all n> N. 


1 We use [a, b] to indicate the closed interval a < x < 6. 



SBC. 7J THE GENERAL THEORY 133 

It is important to note that, in general, the magnitude of N depends not 
only on the choice of e but also on the value of x. 

This last remark may be clarified by considering the series 


Since 


x -f (x ~ l)x -f (x - t)x 2 H {- (x - l)x w ~ 1 H , 

a n (x) + - l)x 2 4 -f (x - l)^” 1 « x n , 


it is evident that 

lim 8 n (x) ££ lim x n * 0, if 0 < x < 1. 

n — * « n a o 


Thus, *(x) — 0 for all values of x in the interval 0 < x < 1, and therefore 
|r n (x)| «* |a„(x) - s(x) | « |x n - 0| * x n . 

Hence, the requirement that |r n (x)| < <, for an arbitrary e, will be satisfied only if 
x n < «. This inequality leads to the condition 

n log x < log e. 

Since log x is negative for x between 0 and 1, it follows that it is necessary to have 

. log « 

n > 

logx 

which clearly shows the dependence of N on both <= and x. In fact, if * = 0.01 and 
x — 0.1, n must be greater than log 0 01 /log 0 1 » — 2/( — 1) - 2, so that N can be 
chosen as any number greater than 2. If € = 001 and x *= 0 5, N must be chosen 
larger than log 0.01 /log 0.5, which is greater than 6 Since the values of log x approach 
zero as x approaches 1, the ratio log */log x will increase indefinitely and it will be im- 
possible to find a single value of N which will serve for e *** 0.01 and for all values of 
x in 0 < x < 1. 


This is the situation which is to be expected in general. In many impor- 
tant cases, however, it is possible to find a single, fixed N, for any preas- 
signed positive e, which will serve for all values of x in the interval. The 
series is then said to be uniformly convergent. 

Definition. The series Zu n (x) is uniformly convergent in the interval 
[a,b] if for each t > 0 there is a number N\ independent of x, such that the 
remainder r n {x) satisfies )r n (x) j < e/or all n > N. 

It is the words in boldface type that give the whole distinction between 
ordinary convergence and uniform convergence. 

To illustrate this distinction in a specific case, we shall discuss the geometric series 

SO 

53 £ n on the interval — J 2 < x < 

n<4 

According to the result of Sec. 1, Example 2, the sum, partial sum, and remainder 
are, respectively, 

J J 

a(x) * 1 S n (x) * ~ > r„(x) « (7-3) . 

I— X I X 1— “X 



134 INFINITE SERIES [CHAP. 2 

The condition |r„(x)| < « gives \x n \ < #(1 - x) or, upon taking the logarithm and 
solving lor n, 

n> log«(I-x) (M) 

log \x\ 


Again it appears that the choice of N depends on both x and e, but in this case it is 
possible to choose an N that will serve for all values of jin ( — HI Given a small €, 
the ratio log «(1 ~ x)/log |x| assumes its maximum value when x ** -\~Vz- Hence if 


N is chosen so that 

N log */ 2 _ log € 

^ log Vi log 2 


then the inequality (7-4) will be satisfied for all n > N. 

Upon recalling the conditions for uniform convergence, we see that the series 2x n 
converges uniformly for — < x < Y- However, the series does not converge uni- 

formly in the interval ( — 1,1), for, in this interval, the ratio appearing in (7-4) will in- 
crease indefinitely as x approaches the values ± 1 . 


Generally speaking, any test for convergence becomes a test for uniform 
convergence provided its conditions are satisfied uniformly, that is, inde- 
pendently of x. For instance, the ratio test takes the form: If there is a 
number r independent of x such that for all large n 

Un+ i(s) 
u n (x) 

then Zu n (x) converges uniformly. Similarly, the comparison test takes 
the form: If 2 v n (x) is a uniformly convergent series such that |w n (x)J < 
v n (t), then 2 u n (:r) converges uniformly. The simplest example of a uni- 
formly convergent series 2 v n (x) is a series of constants. Choosing such a 
series in the comparison test, we are led to the so-called Weierstrass M test : 

Theorem I. If there is a convergent series of constants , 2Af n , such that 
| u n (x) [ < M n for all values of x on [a,b\, then the series 2 u n (x) is uniformly 
(and absolutely) convergent on [a f b\. 

The proof is simple. Since 2 M n is convergent, for any prescribed e > 0 
there is an AT such that 


< r < 1, 


M n+i *T Af n 4.2 + M n + 3 + * * * <6 for all n > N. 

By the ordinary comparison test 2 u n (x) converges for each x, so that r n (x) 
is well defined. We have, moreover, 

|r»(x)| = K+i(x) + u n+ 2 (x ) H 1 < |w„+i(x)| + |u„ +2 (x)| H 

< M n + 1 + M n+ 2 H — • < t 

for all n > N. Since N does not depend on x , this establishes the theorem. 
The other tests for uniform convergence mentioned above are established 
similarly. 



THE GENERAL THEORY 


135 


SBC. 7] 

The fact that the Weierstrass test establishes the absolute convergence, 
as well as the uniform convergence, of a series means that it is applicable 
only to series which converge absolutely. There are other tests that are 
not so restricted, but these tests are more complex. It should be empha- 
sized that a series may converge uniformly but not absolutely, and vice 
versa. 

* sin yix 

To illustrate the use of the M test consider the series — r — Since | sin nx I <1 

n»l ^ 

for all values of x , the convergent series Z 1 /»* will serve as an M series. It follows that 
Z(sin nx)/n 2 is uniformly and absolutely convergent on every interval, no matter how 
large. 

For another example consider the geometric series Zx n . In any interval [—a, a] with 
0 < a < 1 the series of positive constants Za" could be used as an M senes, since 
|x n | < a n on the given interval and since Za" converges 

The importance of uniform convergence rests upon the following 
theorems : 

Theorem II. Let 2u k (x) be a series such that each u k (x) is a continuous 
function of x in the interval [a, 6]. If the series is uniformly convergent in 
[a, 6], then the sum of the series is also a continuous function of x in [a, b]. 
Theorem III. If a series of continuous functions 'EUn(x) converges uni- 
formly to s(x) in [a, 6], then 

rp rP rP rP 

/ s(x)dx~ Ui(x)dx+ u 2 (x) dx 4 h/ u n (r)dx ~ | , 

Ja Ja Jet J a 

where a < a < b and a < (3 < b. Moreover , the convergence is uniform with 
respect to a and (3. 

Theorem IV. Let Xu k (x) be a scries of differentiable functions that con- 
verges to s(x) in [ a,b ]. If the series Zi4(x) converges uniformly in [a, 6], then 
it converges to s'{x). 

The proof is not difficult, and serves well to illustrate the idea of uniform convergence 
(see words in boldface). In Theorem II, if x and x 4* h are on (a, 6], we have 

«(x) * 8 n (x) + r n (x), 

six 4- h) ** s n (x 4- h) 4- r n {x 4* h), 

and hence 

+ h) six) ~ 8nix 4 -h) - 8nix ) + r n (x 4 - h) - r n (x). (7-5) 

Given « > 0, pick n so that |r«(() i < « for all t on [a, b]. Now, s«(x) is a finite sum of 

continuous functions, hence continuous. Therefore 

|«n(x 4 -h) - s n (x) | < « 

whenever |A] is sufficiently small. From (7-5) it follows that 

|*(x 4* h) — a(x)| < |a„(x 4* h) - s«(x)| 4* |r„(x + A)| 4- k«(x)| 

< c 4- « 4* « 



INFINITE SBKIES 


186 


[chap. 2 


This shows that | s(x 4* h) — s(x) | becomes arbitrarily small provided | h | is sufficiently 
small, and hence a(x) is continuous. 

For Theorem III, note that *(x) and r n (x) are continuous by Theorem II. Hence 


J s*{x) dx « J s n (x) dx 4 

hat \r n (x) | < e 

| j s(x) dx — J s n (j) dx j < j J t dx 


fr., 

Jet 


(x) dx. 

If we choose n so large that \r n (x) | < e for all x on [a,b], then 
r& rt* 


|j9 — a | c < (6 — a)e. 


Since the finite sum $ n (.c) can be integrated term by term and aim e (6 — a)t is arbitrarily 
small independently of a and 0, the desired result follows Theorem IV follows from 
Theorem III when u' k {x) is continuous; 1 we simply write down the differentiated series 
and integrate term by term. 

A geometric interpretation of uniform convergence may he obtained by 

considering the graphs of y «= s(x) 
and of the nth approximating curves 
y — s n (x). The condition | (rr) j 
< € is equivalent to 

s(x) — t < s n (x) < s(x) + c (7-6) 

which means that the graph of 
y — s n (x) lies in a strip of width 2c 
centered on the graph of y = s(x) 
(see Fig 4). No matter how narrow 
the strip may he, this condition 
must hold for all sufficiently large n; 
otherwise the convergence is not uniform. 

With such an interpretation, many facts about uniform convergence 
become rather obvious. For example, the conclusion of Theorem III is 



rfi {0 

/ s(x) dx = lim / s n (x) dx 

Ja n —* aa da 


(7-7) 


and the truth of (7-7) is strongly suggested by considering appropriate 
areas in Fig. 4. 

A graphical illustration of nonunijorm convergence is given in Fig. 5, 
Here, the partial sums 




n 2 x 


1 + ft 3 * 2 

are plotted for n = 3, 5, and 10. By inspection of (7-8) 
s(x) ** lim s n (x) *» 0, — oo < x < oo. 


(7-8) 


1 A proof free of this restriction is given in K. Knopp, “Theory and Application a I 

Infinite Series/’ p. 343, Blackie <fe Son, Ltd., Glasgow. 



THE GENEIUL THEORY 


13? 


SEC. 7] 

Nevertheless the approximating curves (7-8) have peaks near x * 0 which 
grow higher with increasing n. Since y = s„(t) does not lie in a strip 
— c < y < € for arbitrarily small 1 € arid all large n, the convergence is not 
unilorm in any interval containing the point x = 0. 

By looking at Fig 5 one cannot easily sec whether the areas under the 
curves y ~ s n (x) tend to 0 or not; that is, one cannot tell whether (7-7) 



holds or not A short calculation based on (7-S) shows that, in fact, (7-7) 
does hold Thus, the coin lusinrt of Tht oreni III nnt\ be true even when 
the convergence is not uinfoiin It i-< lett tor the student to verify that 
(7-7) does not hold when a - 0, 8 ~ 1. and, instead of (7-8), 


6 n (x) = - 


1 4- iru J log n 


(7-9) 


The graphs of y = & H (x) in (7-9) give a figure quite similar to Fig. 5. 


PROBLEMS 

1. rhf» pupal sum- of i senes an s/r) ■* r” Show tint the series is uniformly 
<ouvergml m the mtuvd [(), 1 ' 2 I 

2. By using tin* definition of umfoim <onveiR< me, show that 

1 __ 1 _ 

j + 1 lx + l)tr -f 2) (x 4- n - l)(.r 4 n) 

1 In this caw the condition does not even hold for laige values of #. 



138 


INFINITE SERIES 


{chap. 2 

Is uniformly convergent in the interval 0 < x < 1. Hint: Rewrite the series to show that 
s n (z) «* \/{x -f n) and therefore *»(x) — s(x) * l/(x + n). See Prob. 1* Sec. 1. 

3 . Test the following series for uniform convergence: 

S 2(10*)", Sn(Bin *)», 2 


4. Test for uniform convergence the series obtained by term-by-term differentiation 
of the four series given in Prob. 3. 

5. Plot the sequence s n (x) » nx/( 1 -f nx) versus x for 0 < x < 1 and for n «= 10, 
100, 1,000. Does lim s n (x) ** s(x) exist for every r? Is the convergence uniform on 


0 < 1? 


Is s(x) continuous? Does lim 


J s n (x) dx 



dx for all a, 0 on [0,1]? 


6, If * n (x) * 2nxe~ n **, 0 < x < 1, show that 

lim / s n (x) dx — / lim s„(x) dx « 1. 

n -> <*> J o Jo n 


Is the convergence «»(x) — » s(x) uniform? 


Problem for Review 

7 . Show that 2o« converges absolutely if lim |a«| r < 1. //in/; Choose r' so 
that r < r' < 1. Then V^|a n | < r' for sufficiently large n, and hence |a„| < (r') n . 


POWER SERIES AND TAYLOR’S FORMULA 


8. Properties of Power Series. One of the most important types of 
infinite series is the power series 1 

QO 

]£ a n x n = oq + a x x + a ** 2 H h a n x n (8-1) 

so called because it is arranged in ascending powers of the variable. Typi- 
cal examples are given by the three series 2 

2x n nl, * 2>x n , (8-2) 

n! 


which were already encountered in the foregoing sections. 

For many power series the region of convergence is easily determined by 
means of the ratio test. In the first series (8-2), for instance, the ratio of 
two successive terms leads to 


x n nl 


x n ~ l (n - 1)! 


\xn\ » \x\n 


for x 7 * 0 


1 Throughout Secs. 8 to 14, X means X rather than 2* 

o 1 

1 It is customary to take 0! *» 1, so that the relation n! ** n(n — 1)1 will hold for 
• las well as for n » 2, 3, 4, 



SEC. 8} POWER SERIES AND TAYLOR’S FORMULA 139 

and hence the series converges only for x * 0. In just the same way it is 
found that the second series gives a ratio \x\ /n, which approaches zero. 
Hence the second series converges for all x. The third series is the geo- 
metric series, which, as we know, converges for \x\ <1. 

It is a remarkable fact that every power series, without exception, be- 
haves like one of these three examples. The series converges for x « 0 
only, or it converges for all x, or there is a number r such that 1 the series 
converges whenever \x\ < r but diverges whenever | x | > r. The number 
r is called the radius of convergence , and the interval |x| < r is called the 
interval of convergence. The fact that every power series has an interval of 
convergence may be deduced 2 from the following theorem: 

Theorem I. If I,a n x n converges for a particular value x = x Qf then the 
series converges absolutely whenever |x| < |jr 0 | arid uniformly in the interval 
\x\ < | Xi\ for each fixed Xi such that |x l( f < |x 0 | And if it diverges for 
x = £o, then it diverges for all x such that |u*| > | x 0 1 - 

To establish Theorem I, observe that lim a n Xo = 0, since Xa n XQ con- 
verges (Sec. 1, Theorem 1). Hence | a n xS | <1 for all sufficiently large n, or 


1 

|a n | < for all n > N, say. 

I X 0 i" 


(8-3) 


This shows that 2|a n | |.r| n converges by comparison with the geometric 
series 



provided |j| < |.r 0 |. The statement concerning uniform convergence is 
established by the same calculation, since w( I .r } j ' [i*o|) n serves as an M 
series for the Weierstrass M test. Finally, the statement concerning diver- 
gence follows from the lesult on convergence. That is, if the series con- 
verged for x, it would have to converge for .r ( >, since |x 0 | < \x\, and this 
is contrary to the hypothesis 

The uniform convergence mentioned in Theorem I shows that a power 
series represents a continuous function for all values of x interior to its 
interval of convergence (see Theorem II, Sec. 7) For instance, 2x n = 
1/(1 — x) is continuous for \x\ < 1, though not at x = 1. We shall soon 
see that such functions not only are continuous but have derivatives of all 
orders and the derivatives can be found by termwise differentiation of 
the series. 

1 For simplicity of nomenclature one may incorporate the first two cases into the third 
by allowing r « 0 and r * *>. The case r « 0 arises when the series converges for z « 0 
only, whereas r *= «© if the series converges for all x. 

* A complete discussion is given in Sec. 16. 



140 


INFINITE SERIES 


[CHAP. 2 


As an illustration of this faot consider the geometric series Xx n mentioned 
above. Term-by-term differentiation yields the series 2 nx n ~~ l . Because 
of the coefficient n, which tends to infinity, one might expect the latter 
series to have a smaller interval of convergence than the former. Actually, 
however, the intervals are the same. Since 


nx n ~ l 

; 

1 71 

(n - l)x n ~ 2 


X 

[ n — 1 


| x | , as n — > oo, 


the ratio test shows that the differentiated series, like the original series, 
has the interval of convergence |x| < 1. A similar result is found if we 
differentiate repeatedly. Each differentiation multiplies the ratio by 
n/(n — 1). Inasmuch as n/(n — 1) — * 1, this factor does not change the 
limit of the ratio, hence does not change the interval of convergence. 

For many power series the ratio |a„+j/a n | has no limit as n —► <x>, and 
the foregoing analysis does not apply. However, suppose the series (8-1) 
converges for some value x — x Q ^ 0 , so that, as before, we have the esti- 
mate (8-3). If \x\ < |x 0 |, the differentiated series 2na n x n-1 converges 
by comparison with 



(Note that the latter series was shown to be convergent in the previous 
paragraph.) The same calculation establishes uniform convergence of the 
derivative series if \x\ < |xj| < | x 0 1 , since 


y 


n 



serves as an M series for the Weierstrass M test. Hence, the result of the 
differentiation is actually the derivative of the original series Ha n x n (see 
Theorem IV, Sec. 7). 

The foregoing argument is practically identical with that used to prove 
Theorem I. A third use of the same method establishes the corresponding 
result for the integrated series 2a n x n+1 /( n + I)- In this case the compari- 
son series are, respectively, 


, i /My 

n + 1 \ | Xq j / 


and 


n + 1 \ | x 0 1 / 


Summarizing this discussion we can state the following, which is perhaps 
the most important and useful result in the whole theory' of power series: 

Theorem II. A power series may be differentiated (or integrated) term 
by term in any interval interior to its interval of convergence . The resulting 
series has the same interval of convergence as the original series and represents 
the derimtive (or integral) of the function to which the original series converges . 



141 


SEC. 8] POWER SERIES AND TAYLOFt’s FORMULA 

Consider, for example, the geometric series 

(l-xr'-l+i+r 5 ! + z n 4 , |z| < 1. (8-4) 

Differentiating termwise we obtain 

(1 - x)" 2 *» 1 + 2x + 3x 2 4 f ax"" 1 + • * ■, |x| < 1. (8-5) 

Differentiating again gives an expansion for (1 — x) ~ 3 , and so on Since the series (8-4) 
converges for |x| < I, Theorem 11 shows without further discussion that all these other 
expansions are also valid for |xj <1 

On the other hand, if the series (8-4) is integrated termwise from zero to x, there results 
an expansion 

x 2 x 3 x n 

-log (1 ~ x) * x + ~ + - 4 h i , |x| <1, (8-6) 

z o n 

which can lie used for numerical computation of the logarithm, 

liquations (8-4) to (8-0) give power-series representations for the func- 
tions on the left. It will now be established that such representations are 
always unique 

Theokem III. 7/ hvo power writs conirrgt fo the same sum throughout 
an interval , thm corresponding co* ffn tents are ajual. 

For proof, assume that S«„.r w — so that, by (2-3), 

0 = (do — bo) + («i — l>i)x + (ti 2 — b>).r +■*••+ (u n — b n )x tl + • * *. 

The choice x *= 0 yields a 0 — bo Differentiating with respect to x yields 

0 ~ (ai - bi) + 2(a 2 — b 2 )x H {- n(a n - b n )x n " l H 


and if we now set x — 0, wo get a\ - /q. Fpon differentiating again and 
setting x - 0, wo get a 2 ~ bj , and so on. 

This process not only shows that the coefficients are uniquely determined 
but yields a simple formula for their values Let 

f(x) * a 0 + ciiX + a 2 x z + • • + a n x” 4 , for \x\ < x 0 . 


Upon differentiating n times we get 

f (n) (x) -0+0 + 04 + 0 + //!«„ 4 

where the second group of terms “4 — ■" involves x, x 2 , or higher powers. 
These terms disappear when wo set x = 0, and lienee / fT °(0) - n\a n , or 


/ (r,) ( 0 ) 


(8-7) 


In the following section wo shall be led to the same formula (8-7), though 
by an entirely different method 

The algebraic properties described in See. 2 for series in general give 
corresponding properties for power series: Two power series may be added 
term by term, a power series may be multiplied by a constant, and so on. 



INFINITE SERIES 


142 


(chap. 2 


Since power series converge absolutely in the interval of convergence, The- 
orem IV, Sec. 6, yields the following additional property: 

Theorem IV. Two power series may be multiplied like polynomials for 
values x which are interior to both intervals of convergence. Thus, 


(SOnX W )(S6 n X n ) - XCnX", 


where c n = ao&» + afin^i + 02^—2 + * • * + a n b 0 . 


So far, nothing has been said about the behavior of power series at the ends of the 
interval of convergence. As a matter of fact all behaviors are possible. For example, 
each of the series 





(~-D n * n 

n 


has \x\ < 1 as interval of convergence. However, the first series converges at x « 1 
and —1, the second diverges at x * 1 and —1, the third converges at x — —1 but 
diverges at x « X, and the fourth diverges at x = —1 but converges at x ** 1. 

For applications, the most important theorem concerning the behavior at the ends 
of the convergence interval is Abel’s theorem 1 on continuity of power series, which 
reads as follows: 

Asia’s Theorem. Suppose the power series XanX n converges for x «* xo, where xq may 
be an end point of the interval of convergence. Then 

lim Sa„x n «* 2a n x" 

* *0 


provided x — ♦ Xo through values interior to the interval of convergence 

To illustrate the theorem, let x — * —1 through values greater than —1 in the series 
(8-fi), The limit of the left side is — log 2, since the logarithm is continuous, and the 


limit of the right side is 

l 


(~ir 


by virtue of Abel's theorem. Hence, 


log 2 


1 1 1 (- l)"* 1 

1 -~5 + o-i+‘*- + L — — +•••• 

2 3 4 n 


As another example of Abel’s theorem, let x — ► 1 in Theorem IV to obtain the fol- 
lowing: // 


Cft * acfan -f Q>lbn-~l H h u n 6o, 


then (Za n )(£hn) ■* Xc n provided each series is convergent. Hence, with the particular 
arrangement of the product series which is given by Xc n , we do not need absolute con- 
vergence as in Theorem IV of Bee. 6. 


PROBLEMS 


1. Find the interval of convergence, and determine the behavior at the end points 
of the interval: 


T 'ln r 2n +1 <n. 3 r 2n 

2 (— 



1 A proof is given in I. S. Sokolmkoff, “Advanced Calculus/’ pp. 278-279, McGraw- 
Hill Book Company, Inc., New York, 1939. 



SEC. 9] POWER SERTE8 AND TAYLOR*S FORMUtA 143 

3. Show that the radius of convergence of 2o„x n is given by r ~ lim \a n /a n +\ t 

n oo 

whenever this limit exists. 

3. (a) By letting x ** — / 2 in (8-4), obtain the expansion 

— - 1 -< 2 + * 4 - 1 6 +---+(-\) n t u + •••. 

1 + t l 


(6) By integrating from zero to a:, obtain an expansion for tan 
suit (b), show that 



(c) Using your re- 


4 . (a) Show that the series y « Xx n /n ! satisfies y f * (5) Deduce an expansion for 

e r . For what values of x is this expansion valid 7 fc) Obtain series expansions for e and 
1 /e by taking x » =bl m (b) (d) Using youi sene's, compute e and l/e to three signifi- 

cant figures, and cheek your work by finding the pioduct c(l/e). 

6. Using results given in the text, express the following integials as power series: 


r.A , r 

Jo 1 + t i Jo 


log (1 4 U) dt, 


L 


U dt 

o d~~~ flj*' 


Hint • In the third ease, for example, let x - t z in (8-5), multiply through by f 6 , and 
finally integrate term by term 

6. By multiplication of series obtain the expansion of (1 + .r -f x 2 -f • • * + x n + * * -) 2 . 
In particular, compute the coefficients of 1, x, x 2 , x 3 , and x" in the product series. 


9. Taylor’s Formula. The usefulness oi power series is greatly increased 
by the so-called Taylor formula , which yields the power-series expansion for 
an arbitrary function f(x) together with an expression for the remainder 
after n terms Let /(.r) be a function with a continuous nth derivative 
throughout the interval \a,b]. Taylor’s formula is obtained by integrating 
this nth derivative n times in succession between the limits a and x, where 
x is any point on [a,b]. Thus, 


/> V) dx = = f n ~ u (s) — / (, ‘ _1> (fl) 

Ja I a 

j Z idx) 2 = f X f n ~ l) (x) dx -fj {n ~ l) (a) dx 

= f (n ~ 2 \x) -f n ~ 2 \a) - (x - rt)/ (n_I) ( a ) 

f X f [>>(*) ( dx )* = f (n ~‘ 6) ix) -/ ( ’- 3, (a) - (x- a)f {n ~ 2) (a) 

Ja Ja Ja 

- 

2 ! 


/;••• jj (n \x) (dx) n - f{x) - /(a) - (x - a)f'(a) - (j ~ f"(a) 

(x - a)"~ l 

-f in ~ l> (a). 


in - 1)! 



[chap. 2 


144 INFINITE SERIES 

Solving for f(x) gives 2 

/(*) - /(a) + (* - a)f'(a) + 

2 1 

(x — a )”"" 1 

H 777 f {n ^\a) + R n) (9-1) 

(n — 1)! 

where R n «= f . . . f / (n) (*) (dr) n . (9-2) 

The formula given by (9-1 ) is known as Taylor’s formula, and the particu- 
lar form of R n given in (9-2) is called the integral form of the remainder after 
n terms. The Lagrangian form of the remainder, which is often more use- 
ful, is 

(x — a) n 

Rn = 7-/ (n, (f), a < £ < r. (9-3) 

n! 


To derive this from the form (9-2), let M be the maximum and m the minimum of 
f (n) (l) for a < t < x. Then the integral (9-2) clearly lies between 

f ... f M (dx) n and f ... f m ( dx ) n . 

J a J a J a J a 

Upon carrying out the integration we find that these bounds are 


Or - a) n 
n t 


M and 


(r - a) n 


- m, 


respectively. Since the continuous function assumes all values between its maxi- 

mum M and minimum m , there must be a number t » £ such that (9-3) holds We 
have written our inequalities for the case a < x; in any case, £ is between a and x. 


In general the remainder R n depends on x, as is obvious from the repre- 
sentation (9-2). It may happen, however, that f(x) has derivatives of all 
orders and that the remainder R n approaches zero as n — > 00 for each 
value x on \a,b]. In this case \\e obtain a representation for f(x) as an in- 
finite series 


/(*) 


^ f in) (a)(x - a) n 

n sa*0 ^ • 


(9-4) 


and R n now gives the error which arises when the series is approximated by 
its nth partial sum. The series in (9-4) is called the Taylor scries for f(x) 
about the point x = a. The special case 


m « 


* / (n) ( 0)x B 

h; »! 


(9-5) 


is often called Maclaurin’s series y though Taylor’s work preceded Mac- 
laurin’s. 



SEC. 9] POWER SERIES AND TAYLOR’S FORMULA 


To illustrate* the use of Taylor’s formula, lot /Or) *■» e*. Then f\x) 
* and hence / (w) (0) ** 3 Equation (0-5) suggests that 


e x 


r L x* x” 


145 

f"(x) * «* 

(0-0) 


and indeed, bv the ratio test this series converges for all x, However, to show that it 
converges to c x we must consider the remainder, which takes the form 


Rn 



0 < t < T, 


(9-7) 


when we use (9-3). Since this approaches zero 1 as n -* <*, the aeries docs converge to e*. 

As anot her example we find the expansion of cos x m powers of x - (r/2). The values 
of /, f", are, respectively, 


cos x, — s in t, —cos x , sin x. (9-8) 

Sime the next term is / 1V ** cost, the next tour derivatives repeat the sequence (9-8), 
the next four repeat again, and so on Evaluating at x — tt, 2 w e get, respectively, 

0, -1, 0, 1, 0, -1, 0, l; 0, —1,0, 1 , . . 


and hence Kq. (9-4) suggests the expansion 




<*W*) 


To determine if the senes conveiges to the Junction on the left, we consider the re- 
mainder after n terms Now, (0-8) gives f yn) (x) — -t mm x or ri cost, so that (9-3) 
implies 1 R n | < u — it 2| w , n ’ Since bm R n - 0, the expansion (9-9) is valid 

Upon setting t ** ir y 2 — t and noting that ros(jr 2 — /) — s m/, wm get an expan- 


sion lor sin £ 


sin t * t 


t* r t 1 

— — — T 

3* .V 7* 


* 4- - - * + • 

(2a -h 1)' 


(9-10; 


which is consistent with (9-5) It is left as an exercise lor the reader to obtain a similar 
expansion for the cosine by us* of (9-5) and (9-3). 


cos t = 1 




+ 


(2«) » “ + 


(9-11) 


In these examples the fact that the series converges to the function was 
established by direct examination of R n . Such examination is necessary 
even when the series is found to be convergent by other means. For ex- 
ample, if we define 

rx) = m = o , 


it (‘an be show n that the Taylor series about x = 0 converges for all x but 
converges to /Or) only when x = 0. The trouble with this function is that 
it does not admit any power-series expansion valid over an interval con- 
taining x - 0, and we have the following: 

1 The fact that (9-6) converges shows that x n /nl — * 0. (Cf. Prob. 5, Sec. 5.) 



146 


INFINITE SERIES 


[CHAP. 2 

Theorem I. Suppose a function f(z) admits a series representation in 
powers of x — a, so that f(x) * Xa n (x — a) n for some interval |x — a\ < c. 
Then the Taylor series generated by f(x) coincides with the given expansion 
[and hence, the Taylor series converges to f(x)]. 

For proof, differentiate 1 n times and set x * a, just as in the discussion 
of (8-7). It will be found that a n — f in) {a)/n \ , and hence the given series 
is identical with the Taylor series. 

Theorem I shows that a valid power-series expansion obtained by any 
method whatever must coincide with the Taylor series. For instance, to 
find the Taylor series for sin x 2 about x = 0, we set t =» x 2 in (9-10). This 
is far simpler than direct use of Taylor's formula, as the reader can verify. 


Example: Obtain the expansion of 


fix) 


1 

tx - 2 )(x - 3) 


in powers of x — 1. 

With t » x — 1, the given function becomes 


1 -1 1 111 
(/ - 1)(! - 2) ” t - l + t - 2 " X - < “ 2 1 - Mi 

(9 - ,2 > 

when we use partial fractions and the known formula for sum of a geometric series. 
Upon recalling that t «* x — 1, we get the required result 

IW3) 

Since the two geometric series (9-12) converge for \t\ <1 and |<| <2, respectively, 
the expansion (9-13) is valid for \x — 1 1 < 1. By Theorem I, this expansion coincides 
with the Taylor series. 

PROBLEMS 


1. For the following functions find the Taylor series about the point x ** 0 and also 
about the point x — 1: 

e 2z t sin vx t cos (x — 1), 2 -f x 2 , (x -f* 2)~ l . 

2 . (a) Expand e x about the point x * a by writing e x * e a e x ~ a and using (9-6). 
(6) Expand logx about x 1 by writing logx *** log [1 — (1 — x)] and using (8-6). 
(c) Obtain the general Taylor series from Maclaurin’s. Hint • If g{t) =» f(a -j- t), then 

f(a ~M) - ?(*) ~ S^ (n) (0)r/n! - Xf^ n) {a)t n /nl 


Now let < « x — a. 

1 The fact that the series now considered are in powers of x — a rather than x causes 
Ho trouble. By a simple translation of axes, £ » x — a, these series become power series 
of the type considered in the preceding section, hence are subject to the theorems of the 
preceding section. 



147 


SEC. 10] POWER SERIES AND TAYLOR’S FORMULA 

3. Expand the following fractions about the point x ** 1: 

1 1__ I 

x x 2 — 4* x(x 2 — 4) 

4 . Show that the Taylor series for sin x in powers of x — a converges to sin x for every 
value of x and o, and find the expansion of sin x in powers of x — tt/6. 

6. Obtain the Maclaurin aeries for cos x by differentiating the series for sin x. 

6. By means of the known series for e u , sin u, and log (1 — u), find Taylor's expansions 
for 

e~ x \ sin x 2 , e x -f e~* x , e x — e“* x~ 2 log (1 -j- x 4 ). 

7. What is the Taylor series for (I 4* x) p if p is constant? Find the interval of con- 
vergence, and discuss the absolute convergence at the end points of the interval. Hint: 
Use Theorem IIfc, Sec. 5. Analysis of the remainder R n is difficult and may be omitted, 
A proof that the series converges to (1 -f x) p will be found m Sec. 12. 


10. The Expression of Integrals as Infinite Series. Many difficult in- 
tegrals can be represented as power series. For example, it* we let x = t 2 
in the series (9-11) for cos x , we get 


cos r 


t A t* 

1 + - + ■ 

2! 4! 


and hence, integrating term by term, 


/; 


cos t 2 dt = x 


5-2' 


(-l)V 

•+• 1 -- 

(2 n ! 


+ ••■ + 


(-l)V 


n „4n + l 


(4 n + 1 )(2n)! 


+ • 


( 10 - 1 ) 


This integral is called the Fresnel cosine integral; it is important in the theory 
of diffraction. Although the Fresnel integral is not expressible in terms of 
elementary functions in closed form, the expansion (10-1) is valid for all x 
and gives a representation which is entirely adequate for many purposes 
Sometimes one may obtain a power series involving a parameter rather 
than the variable of integration as in the last example. To illustrate this 
possibility we shall express the arc length of an ellipse as a power series in 
the eccentricity /c. If the equation of the ellipse is given in parametric form 
as 

x = a sin 0, y — b cos 6 , a > 6, 


then the arc s satisfies 


ds 2 = dx 2 + dy 2 — (a 2 cos 2 6 + b 2 sin 2 0) d$ 2 . 

Upon noting that cos 2 0=1— sin 2 0, we obtain 
ds — a VT^T 2 sin 2 $ d$, 

where k «* (a 2 — b 2 ) **/ a is the eccentricity. Hence, the arc from 0 « 0 to 
6 « 4>is 


8 == a f Vi — k 2 sin 2 0 dd ss aE(k 

Jo 



148 


INFINITE SERIES 


[CHAP* 2 

The integral E(k,<t>) defined by this equation is called the elliptic integral of 
the second kind . Although E(k,<t>) is not elementary, it may be expressed as 
a power series. 

By the binomial theorem (Sec. 12) 

(1 - k 2 sin 2 e)K » 1 - ]/ 2 k 2 sin 2 B - y s k 4 sin 4 B (10-2) 

for k 2 < 1, which is the case when 6 ^ 0. Since 

i + yk 2 + y$k 4 + • • • 

serves as an M series, (10-2) is uniformly convergent and term-by-term 
integration is permissible. Hence we obtain the desired expression 

1 „ ft n k 4 ft 

E(k,<p) « <f> k 2 sin 2 6 dO / sin 4 6 dd — * • • 

2 •'o 2-4 A) 


1 -3-5 ... (2n - 3) 
2-4-6 . . . 2n 



In a Bimilar manner it can be shown that the elliptic integral of the first kind (ef 
Example 3, Sec. 20, Chap 1), 


lias the expansion 




1-3 f* 

sin 2 0 dd -} k 4 / sin 4 6 do 4* • 

2 4 Jq 


The elliptic integral of the third kind is 

ft 


+ r~~ „ — — sm “" 0(19 H • 


1-3 5 ... (2n - 1) 
2-4~6T7.2n" 


U(n,k,4>) 


r — 

h (1 4- n 


dO 


sm 2 0)\/l — k 2 sin 2 0 


and this, too, can be expressed as a series by expanding the radical. 

Any integral of the form 

J (a sin x + b cos x + c)^ dx 

or of the form 

f i?(i,Va^ f &r 3 -f cx 2 + dx + e ) dx, 1 R(x,y) = rational function, 

is expressible in terms of the elliptic integrals 1 together with elementary 
functions. For this reason elliptic integrals have great practical impor- 
tance and have been extensively tabulated. 

1 See, for example, P. Franklin, “Methods of Advanced Calculus/’ chap. 7, McGraw- 
Hill Book Company, Inc,, New York, 1944. 



sue. Ill 


POWER SERIES AND TAYLORS FORMULA 


149 


PROBLEMS 


- dt. 


1. Expand the following integrals as power series: 

dt, f sin (t 2 ) dt, — t 

Jo Jo t Jo Jo t 

2. Express / e x * iat dt as a powe r series in x. Hint By Wallis' formula, 

Jo 

I sin” i 

Jo 


1 1 dt 


(n - ])(w — 3) 


2 or 1 


n(n — 2) ... 2 or 1 


where <x ** 2 if n is odd and a * ir if n is even. 
3. Express the incomplete gamma Janet ion 


f P“V-* 
Jo 


dt 


as a series in powers of x. For w hat values of x and p is your expansion valid? 
4. The beta Junction is defined by 


B(p 


} q) m f'x»-\ 1 - 
Jo 


I)*- 1 dx. 


Express this a& a scries by using the binomial theorem for (1 — x) H ' 1 and integrating 
term by term Fur whul values of p and g is the resulting series absolutely convergent? 
(See Theorem 116, See 5. Although the range of integration includes the value x «* 1, 
w Inch in an t ud point of t he convergence interval, the integration is easily justified. Thus, 

one might consider / and let x -* 1 through values less than 1. The desired result 
Jo 

then follows fiom Abels theorem, Sec. 8 ) 


11. Approximation by Means of Taylor’s Formula. If a function f(x) 
has a convergent Taylor series, then the partial sums of that series can be 
used to approximate the function. In this way, calculations of great in- 
trinsic complexity are reduced to calcuJat ions mvohing polynomials. The 
method is especially important because Taylor’s formula not only gives a 
polynomial approximation but gives a moans of estimating the error. 
Thus, the remainder R n in (9-2) and (9-3) is precisely the difference be- 
tween f(x) and the nth partial sum of its Taylor series. 

To illustrate the use of Taylor's series for numerical computation, let 
us find sin 10° within =bl0~ 7 The value 10° -= tt/ 18 radian is closer to 
zero than to any other value of x for which sin x and its derivatives are 
easily found, and hence the expansion is taken about the point x = 0. To 
estimate the number of terms required, (9-3) gives 


l«n| 


/ <n) «) „ *’ 
— X < — 

n! n ! 



(0.175) n 


n! 


(1W) 


when a — 0, and when we set x = x/18 = 0.175 and recall that sin x to- 
gether with its derivatives is lass than 1 in magnitude. The successive 



150 


INFINITE SERIES 


[CHAP. 2 

bounds for R n as given by (11-1) may be computed recursively; indeed, the 
nth value is obtained by applying a factor (0.175/n) to the preceding one. 
For n * 1 the bound (11-1) is 0.175 and the next few are as follows: 


Value of n 

2 

3 

4 

5 

6 

Bound for \R n \ 

1.5 X 10~* 

8.8 X 10 -4 

3.9 X 10~ 6 

1.4 X 10-« 

4.0 X 10~ 8 


From a list such as this the n sufficient for any prescribed accuracy can be 
determined at once. In particular, an accuracy of ±10“ 7 is found if we 
take n » 6. Thus 


sm 


TT / 7T \ 3 1 / it \ 6 1 

10 ~ 18 ~ (.18/ 3! + \Ts) 5! + Re ’ 


where | R$ | < 4.0 X 10~ 8 ; more explicitly, 0 > R& > —4.0 X 10~ 8 , since 
/ <6) (£) =k — sin $ < 0. Inasmuch as the next term of the series is zero, 
the first six terms are the same as the first seven terms. Hence the error is 
also equal to R 7f where 

0>R 7 > —1.0 X 10~ e . (11-2) 

An improvement of accuracy such as this is to be expected whenever the 
series is terminated just before one or more terms with zero coefficients. 

In modern computing practice an automatic computing machine is so programmed 
that it keeps track of the remainder, which can often be estimated recursively as in 
this example. The machine is then instructed to take as many terms as are needed to 
make the remainder less than some preassigned amount. This process was illustrated 
in the foregoing calculation, where the value n ** G was chosen, not at random, but by 
consideration of the desired accuracy. 

The reader may have noticed that the series for sin (ir/18) is an alternating senes with 
terms decreasing in magnitude. Hence, the estimate for the error (11-2) could have 
been found by Theorem II of Sec. 6. Taylor s formula, however, has the merit of apply- 
ing to general power series, whether alternating or not. 

Many important approximations are obtained by using the first few 
terms of the Taylor-series expansion instead of the function itself. For ex- 
ample, the formula 

k - y"l 1 + (y') 2 r H 


for curvature of the curve whose equation is y - f(x) yields 


k *= 



3 „ 13 5 

- ( y ') + (v ') 4 

2 Ky 2! 2 2 U 



when we use the binomial theorem. The first-term approximation k ££ y" is 
sufficient for most applications. 

As another example, in railroad surveying it is frequently useful to know 
the difference between the length of a circular arc and the length of the 




POWER SERIES AND TAYLOR’S FORMULA 


151 


SBC. 11] 

corresponding chord. Lei r be the radius of curvature of the arc AB 
(Fig. 6), and let a be the angle intercepted by the arc. Then, if s is the 
length of the arc AB and c is the length of the 
chord AB, s * ra and c = 2r sin Since 


sin x * x 1 — cos £, 

3! 5! 

where 0 < £ < x, the error in using only the first 
two terms of the expansion is certainly less than 
x 5 /51. Then, 


c = 2 r sin - 


with an error less than 


2r 


/ a or \ 

\2 ~ 8 - 7 )/ 



2r 


/ a* \ roA 
\32 * 1 20/ ~ 1 , 920 
Therefore, $ — c = a?r / 24 with an error that is less than ra 5 / 1,920. 

Example 1 . For the nonelementary integral / c u * du obtain a polynomial approxima- 


tion valid within -i-0 00001 when 0 < x < 
According to (9-6) and (9-7), 

i 2 x n ~* 1 

€ 


1 +X+~ +•••+7— 0 < { < X. 
2» (n - 1)! n! 


If we set x * u 2 , this becomes 


c u2 - 1 + u 2 + ~ + •••+: — + 0 < { < u 2 , 

2! (n — 1) ! n! 

and integrating from 0 to x yields 


/> 

Jo 


du ** x 4 1 1 1 

3 5 2! X (2 n - l)(n 


f * u 2n 

+ Jo ~nJ 


du. 


To estimate the integral on the right we note that 

e* < e tt * < e 3 * 2 , 

since £ < it 2 and u < x. Hence 


It follows that if we write 


f u ln et du < e xi [ u 2n du 
Jo Jo 

«e 

f « u2 du ^ 22 7~ — ' 

Jo «**i (2n — 


r 2rt-fl 


2 » + 1 


1 )(n - 1)! 


then the error Ran after the term x 2n 1 satisfies 

p x a 

0 < Ran < 


n!(2n + 1) 



152 


INFINITE SERIES 


[chap. 2 


For ah approximation valid within dtO.OOOOl when 0 x <J we choose n large 
enough to make 

- -7 ~~ z < 0 . 00002 . 
n!(2» + 1) 

Since e* < 1.3, the above condition is satisfied when 

n!(2n + l)2 2n+1 > 65,000. 


By trial we find that n «* 4 suffices. This choice of n yields an approximation wffiich 
is too small by 0.00002 at most. If 0.00001 is added to the approximation, we get 



* + ? + £ + S +0 00001 


(U-3) 


within dbO.00001 when 0 < x < H- 

Example 2. Obtain a polynomial approximation for / e* 1 u * dx valid near x «* 0. 

Jo 

Keeping terms as far as x z and no terms beyond x 3 , we have 

-* + (— T>+i(— T>’+- 


Hence 


1 +* + -x 2 + 0-z s +- 


/V** 

Jo 


dx 


X + -+-+. 


where the terms omitted involve x 6 or higher powers. 

This calculation of the series for e Biux illustrates a principle which is often useful. 
Let f(y) * Xb n y n and y = '£a n x n be power series w r ith nonzero radii of convergence. 
If y *» 0 when x ** 0, then the power senes for f(y) as a function of x also has a nonzero 
radius of convergence. This series may be found by substituting the series for y into the 
series for f(y) and collecting terms. 1 * By uniqueness, the series so obtained is the Taylor 
series. 


PROBLEMS 


1. It is desired to approximate a function f{x) by a polynomial p{x), 


p(x) * oq + aix -} h anz n . 


in such a way that at the origin p(x) has the same value and the same first n derivatives 
as f{x). (a) How should the coefficients be determined? Hint: Oq «* p( 0) * /( 0), ai «* 

p f ( 0) « /'( 0), 2a2 *■ p"(0) * /"(0), (6) If the coefficients are determined as m (a), 

what relation does p(x) have to the Maclaurin series for /(x)? 

2 . For the following functions obtain a polynomial approximation valid near x » 0 
by finding the first three nonzero terms of the Maclaurin series: 


tan x, e 1 * 0 * sec x t 


e* sin x 
1 -f e*’ e* ~ l' 


1 For proof, see K. Knopp, “Theory and Application of Infinite Series, 5 ' p. 180, Blackie 

k Son, Ltd., Glasgow, 1928. 



163 


SEC. 12] POWER SERIES AND DIFFERENTIAL EQUATIONS 

3. (a) By means of series compute x — 10 sin (x/10) to three significant figures when 
x ** 1.000. (6) Attempt the same calculation by using a table of sin x (note that a; is in 
radians). How many significant figures for sin x are needed? 

4. If y « 10 (tan x — x)/x % , (a) use series to evaluate y near x *» 0. In particular, 
what is the limit of y as x — ► 0? (ft) Plot y versus x for 0 < x < 0.2. (c) Discuss the 
construction of such a graph by use of a table of tan x r without series. 

6. By use of series compute to three places of decimals: 

(a) e 1 1 - ee 0,1 * 2.7183e° l ; (6) coaJ0° * cos Or/18); 

(r) sin 33° - sin (30° + 3°); (d) ^35 * 2(1 + J$ 2 ) w . 


6. Evaluate by series the first three integrals to three places of decimals and the last 
to two places: 

sin x dx f° A log (1 - z) 


[ sin (x 2 ) dx } [ 
Jo Jo 


j 

Jo 


dz , 


VT — x a Jo 
J (2 — cos x)~ H dx ** J ^1+2 sin 2 dx. 


7. Determine the magnitude of a if the error in the approximation sin a £* at is not 
to exceed 1 per cent. Hint ( a — sin a) fa = 0.01 and sin a = a — (* 3 /3t) + (aVfit) - 


8. Discuss the percentage error in the approximation (1 1-3) as x 0. How would 
the percentage error behave if the term 0.00001 had not been added? (Tins shows that 
it may be better not to alter the Taylor series even when such alteration reduces the 
absolute error.) 

Ste As in Example 1 of the text, obtain a polynomial approximation for the Fresnel 


sine integral 



dt which is valid within ±0.00001 for 0 < x < 3^. 


POWER SERIES AND DIFFERENTIAL EQUATIONS 

12. First-order Equations. One of the most important uses of power 
series is in the solution of differential equations. P'or example, to solve the 
equation y f — y assume that 

y = a 0 + a { x + a 2 x 2 + a 3 x 3 H f a n x n H . 

Then, according to Theorem II, Sec. 8, 

y' = ai + 2 a 2 x + 3 a 3 x 2 + ia 4 x 3 H b (?i + l)a n+1 z n d . 

Since y f = y, Theorem III of Sec. 8 shows that the series for y* and the 
series for y must have the same coefficients. Thus, 

£Zj = dp, 2d 2 ~ 3ci3 ~ d 2 , • » • i (ft d~ ^ . . .. 

Starting with a 0 = c, a constant, we solve for aj, a 2 , ... in succession to 
obtain 

ai-c, a 2 = -. a 3 = ^’ •••» a " = nl' 



154 INFINITE SERIES [CHAP. 2 

and hence 

CX 2 CX 3 €X n 

„. c + car + _ + _ + . 

This discussion is tentative only, since at first we had no assurance that 
the equation y f = y possesses a solution expressible as a power series. 
However, the ratio test shows that the series obtained converges for all x. 
Hence by Theorem II of Sec. 8 the term-by-term differentiation is justified 
and the equation y ' « y has actually been solved. 

As another illustration, we obtain a power-series solution for the differ- 
ential equation 

(1 + x)y' * py, p « const, (12-1) 

such that y « 1 when x — 0. If y « 2a n x n , then 1 

y’ ** ai + 2a 2 x + 3a 3 x 2 -| f (n + l)a n+ |X n H 

xy' * aix + 202X 2 H h na„x n H 

py « poo + paxx + pa 2 x 3 H b pa n x n H 

and the substitution in (12-1) yields 

a\ + (2 a 2 + ai)x + (3a 3 + 2a 2 )x 2 + • • • + [(w + l)a n+ i + na n ]x n + • * • 

« pa 0 + P^i* + pa 2 x 2 H b pa n x n H . 

Equating coefficients of like powers of x gives the set of equations 

a \ ~ poo 
2 02 + 01 = P«i 
3a 3 + 2a 2 = pa 2 


(n + 1 )a n +i + na n * pa n 
which must be solved for the as. 

Since y « 1 at x = 0, we must have a 0 = 1. Then we get, in succession, 

p(p - 1) P(P — l)(p — 2) 

ai *= p, a 2 — * a 3 = » ... 

y 21 3! 

so that the solution is 


y - 1 -f px + 


P(P ~ 1) _ 2 


2! 


* +• 


p(p - l)(p - 2) . . . (p - » + 1) 

H : x n + • 

n! 


(12-2) 


Throughout Secs. 12 to 14 2 means 22- -4 brief review of this sigma notation is given 
in Sec. 2. 



POWER, SERIES AND DIFFERENTIAL EQUATIONS 


155 


SEC. 13] 

Hence, if the differential equation (12-1) has a series solution, that solu- 
tion must be (12-2). However, the ratio test shows that the series (12-2) 
converges for \x\ <1. Hence for |x| <1 the term-by-term differentia- 
tion was justified, and this shows without further discussion that (12-2) is 
a solution for \x\ <1. 

Equation (12-1) can be solved by elementary methods as follows. Sep- 
arating variables, we get 

dy dx 

— = V » 

V 1 + x 

so that log y * p log (1 x) = log (1 + x) p when we recall that y * 1 at 
x = 0. Hence 

y » (l + z) p . 

Comparing this solution with that found formerly 1 gives 


(1 + x)P = 1 + £ 


(p - n + 1) 


which the reader will recognize is the binomial theorem. Since no assump- 
tion was made about p the result ts valid for all p provided \x\ <1. 


PROBLEMS 

1. Obtain power-series solutions of the following differential equations, whidi satisfy 
y 1 at x “0* 

y' =** 2y, xj « y -f x, y' -b y « 1. 

2. Obtain the first three terms of a series solution y * 2a n x" for the problem y ' * 2 esy, 
y 1 when x •» 0. From these three terms compute the curvature k =» y"\\ -J- (y') 2 

at x =» 0, Is your value for curvature exact or only approximate? 

3. By considering the equation y' ** (I — x 2 ) - ^ obtain a series expansion for sin” 1 x. 
In particular, show that 

6 * 2 + 2 3 *2* + 2-4 5-2* + 2*4~fi7-2 7 4 ’ 

13. Second-order Equations. Legendre Functions. To illustrate the 
use of series for solving second-order differential equations consider the 


equation y „ _ xyl + y _ Q (13-1) 

If y = Sa n x n , then y f * 2na n x n ~ l and y" = 2n(n - l)a„x n ~ 2 and, hence, 
y" =s 2a2 + 3 • 2a 3 x + * * * + (n + 2)(n + \)a n + 2 X n + * • * 

— xy * ... na n x — * * * 

y *» ao + a t x -f b &nZ n + * • • • 

1 By Chap, 1 , Sec, 17, the problem has only one solution. 



1S6 


INFINITE SERIES 


[COUP. 2 

According to ( 13 - 1 ) the sum of these three power series is zero, and there- 
fore, by Theorem III of Sec. 8, the coefficient of x n in this sum is zero for 
each n : 

2 a a + ao « 0, coefficient of x°, 

3 -2a3 — a\ + cq = 0, coefficient of x l , 


(n + 2)(n + 1 )a n + 2 — na n + a n « 0, coefficient of x n . 

Hence, 

n+2 (n + l)(n + 2) 

This recursion formula gives, in succession, 


(13-2) 


1 

1 

1! 

t <N 

I 

II 

" 21 001 

1 

1 

~ 3^4 02 ~ 

--a°, 

3 

3 

= — a 4 = 
5-6 

"Si 00 ’ 


®2n 


1 - 3-5 ... ( 2 n - 3 ) 
(2n)I 




n > 3 . 


Similarly, 03 = 0 * cq = 0, 05 = 0 , a 7 — 0, and so on. Hence the solution is 

/ 1 2 1 4 1-3 e i* 3 * 5 8 \ , 

y * do ( 1 x * x x ar — • • • ) + aiX 

V 2 ! 4 ! 6! 8! / 


with the two arbitrary constants oo and a x . There should be two constants 
in the general solution because the equation is of the second order (Chap. 1, 
Sec. 21). 

The ratio of two successive nonzero terms of the infinite series satisfies 


02***" 


2 n - 1 

(2n + l)(2n + 2) 


1 **! 



(13-3) 


as is seen by using (13-2) with 2 n written in place of n. Since the limit of (13-3) is zero, 
the series converges for all x. Hence the term-by-term differentiation is justified, and 
we really do get a solution. 

Upon choosing oq * 0, a% * 1 we see that a particular solution is 


!/i « * 


( 13 - 4 ) 



SBC. 13] POWER SERIES AND DIFFERENTIAL EQUATIONS 157 


and the choice oo ** 1, ** 0 shows that another particular solution is 


y% 


1 -~x 2 
2! 




1*3*5 

8! 


(13-5) 


According to Chap. 1, Sec. 21, two solutions y x and y 2 such as this are 
linearly independent if the equation c x y x + c 2 y 2 25 0 can hold for constants 
c x and c 2 only when c x = c 2 = 0. The solutions (13-4) and (13-5) are in- 
dependent, since one is an odd function and the other is even. 1 Hence the 
expression a Q y\ + a x y 2 is the general solution of (13-1) (Chap. 1, Sec, 21). 
The independence of y x and y 2 can also be deduced from the following 
theorem : 

Theorem I. Let y x ~ Za n x n and y 2 — %b n .r n be power series with non- 
zero radii of convergence, and suppose that y x & 0. If y x and y 2 are linearly 
dependent , then there is a constant c such that b n = ca n for all values of n. 

From ciSa*#” + <' 2 2h n x n » 0, it follows that 


cia n + rd>n 3=3 0, n 0, 1, 2, . . .. (13-6) 

If C 2 °* 0, then (13-6) gives a n * 0 for all n, contrary to hypothesis. Hence eg 9 * 0 t 
and (13-6) gives b„ • (— c\/cl)a n ~ This is the required result. 


Obviously not every differential equation has solutions that can be 
represented by the power series. 2 The following theorem due to Fuchs, 
which we state without proof, 3 gives sufficient conditions for the existence 
of power-series solutions of second-order linear equations. 

Theorem II. Let y" + fi(x)y f + f 2 (x)y - 0 have coefficients f x (x) and 
f 2 (x) which can be expanded in convergent power series for \x\ < r. Th$n 
every solution y can be expanded in a convergent power series for \x\ < r. 

The solution series converges at least for | x | < r but may converge in a larger inter- 
val. For example, consider the equation 

(2 - x)y" + (x - W - V - 0. (13-7) 


Writing (13-7) in the form 



0 


(so that the coefficient of y " is 1 as in Theorem II) we get 


h(x) 


x — 1 1 x — 1 

2 - x “ 2 l-^i’ 


1 11 

h(x) = - YZ~ " 21 -Hr' 

1 A function f{x) is odd if /(— x) « —/(x), ewn if f(—x) ** f(x). 

* Thus, xy f » 1 has a solution y log x which cannot be expanded in Maclaurin's 
series. 

8 A simple proof is given in H. T. H. Piaggio, “An Elementary Treatise on Differential 
Equations and Their Applications ,’ * 2d ed., George Bell & Sons, Ltd., London, lf)2S. 



INFINITE SERIES 


158 


[chap. 2 


Since fi(x) and f%{x) have power-series expansions for |x| <2, Theorem II asserts that 
the solution y also has a power-series expansion valid for |x| <2. Actually, the solu- 
tion of (13-7) is 

x n 

y *■ -f- — x) «■ Ci2 — - -f- ca( 1 — z) t 

nl 

which converges for all x. 


As another example we shall find the complete solution of Legendre's 


equation 


(1 — x 2 )y" - 2 xy' + p(p + 1 )y * 0, 


03 - 8 ) 


where p is a constant. 

By Theorem II this equation has a power-series solution valid at least 
for | x | < 1 . Assuming y = Sa n x n , we get 

y n = Sa n+2 (w + 2) (ft + l)x n 
—x z y" » 2 — a»n(n — l)z n 


—2xy' = 2 — 2na u x n 
p(p + 1)2/ = Sa n p(p + l)x n . 

By (13-8) the sum of these series is zero. Considering the coefficient of 
yields 


<*n+ 2 (n + 2 )(n + 1) » a n [n(n -f 1) - p(p + 1)] (13-9) 

after slight simplification. For all n > 0 we have 

(P - n)(p + n + 1) 

a„ +2 - — a n - — — ~r ’ (13-10) 

(n -f l)(n + 2) 

after factoring the bracket in (13-9) and dividing by (n + 2)(n + 1). The 
coefficients for even n are determined from a 0} and those for odd n from a\. 
Computing the coefficients successively, we get the final result 


a 0 


1 - 


p(p + 1) 

2! 


* 2 + 


p(p - 2)(p j- l)(p + 3) 
4! 




+ 


ai 


(p-I)(p + 2) (p — l)(p — 3)(p + 2)(p + 4) 


3! 


■x* + 


5! 


]■ 


Theorem II guarantees that the general solution can be obtained in this fashion, and 
indeed, by Theorem I, the coefficients of do and ai in the foregoing expressions are in- 
dependent. Equation (13-9) shows that the series converge for |x| <1 when we apply 
the ratio test to the ratio of successive nonzero terms. When p is a positive even integer, 
however, the expression involving ao reduces to a polynomial, hence converges for all x. 
If Co is so chosen that the polynomial has the value unity when x =* 1, we get the sequence 


PM 

PM 


1, 

3 2 1 
2 2 

7*5 4 o 5 ' 3 2 

— x 4 — 2 — X 2 

4-2 4-2 


+ 


3*2 

iY 



POWER SERIES AND DIFFERENTIAL EQUATIONS 


SEC. 14] 


159 


and so on. Similarly, when p is a positive odd integer, the coefficient of a\ terminates 
and we get a sequence of which the first three terms are 

Pi(z) « x, 

n , x 5 - 3 

P z (x) - - X Z - - X, 


P*(x) 


2 

9*7 


2 


_ 2 — X 3 

4-2 4*2 


5*3 

4 * 2 ; 


These polynomials are known as Legendre polynomials; they arise in several branches of 
applied mathematics. 


PROBLEMS 


1. Solve by means of power series: 

v" + y “ o, y" -f y 1, ?y" - xy « 1. 

2. For Kq. (13-7), show that y * 2)a n x n leads to 

2(?i -f- 2)(/i 4* l)a n j 2 ™ (n 4 l)4i„ fj 4- (« — l)a« ** 0. 


(a) By taking a« « » 1, find a solution satisfying y = y f ® 1 at x » 0. (6) By an- 

other choice of ao and a\ find a solution satisfying y ** — ], y' = I at x * 0. 

3. Solve by means of power senes if y( 0) — y'{ 0): 

(x* - 3x + 2)y" + (x 2 -2x- W + (x - 3 )y - 0. 

4. Solve y" — (x — 2)y =0 by assuming y ** £a«(x — 2) n . Also obtain the first 
three terms of the solution m the form y = Za n i n 

5. It can be shown 1 that 


(1-2 xh -f h 2 r H - PoCri + Piix)h 4- Pz(x)h 2 4 b P w (r)A« + • • • . 

V(‘rify thus equality through the terms in h b Hint Expand [1 — (2hr — by the 

binomial theorem, and collect powers of h The function (1 — 2xh 4" /r)“ Va is calfed 
the generating function of t lie sequence P n {x ). 

6. Verify Rodrigues' formula* 

e.(*) - ~~ - D B 

for n «• 0, 1, 2, 3. 


2"n l dx n 


14. Generalized Power Series. Bessel’s Equation. An important dif- 
ferential equation was encountered by the German astronomer and mathe- 
matician F. W. Bessel in a study of planetary motion. The so-called Bessel 
functions which arise from the solution of this equation are indispensable 
in the study of vibration of chains, propagation of electric currents in 
cylindrical conductors, heat flow in cylinders, vibration of circular mem- 
branes, and many other problems of applied mathematics. 

Bessel’s equation is 

x 2 y" + xy' + ( x 2 - p 2 )?/ = 0, (14-1) 

1 E. J. Whittaker and G. N. Watson, 14 Modern Analysis,” pp. 302-303. Cambridge 
University Press, London, 1952. 

*Ibid. 



INFINITE SERIES 


160 


[chap. 2 


where p is a constant. Theorem II of Sec. 13 does not apply to this equa- 
tion, since 

1 p 2 

/iW - - /*(*) » 1 

x ar 

and these functions cannot be expanded in power series near x — 0. For 
this reason we do not expect a power-series solution y = 2a n x n . It was 
shown by Fuchs, however, that a wide class of equations, including (14-1), 
have solutions of the form 1 

00 

V = x p 2a„x n - £ a»x n+p , a 0 * 0 (14-2) 

where p is constant. The theorem of Fuchs reads as follows: 

Theorem I. Let xfi (x) and x 2 f 2 (x) have power-series expansions valid for 
| x j < r. Then the equation 

y" + fi(x)y’ + f 2 {x)y « 0 

has a solution of form (14-2), also valid for \x\ < r. 

Since Bessel's equation gives 

xfi(x) = 1, x 2 / 2 (x) = x 2 - p 2 , 

Theorem I asserts that the series (14-2), when found, will be valid for all x. 
To obtain this series, note that 

X 2 L ««* n+p = i a„x n+p+2 - £ a n _ 2 x” +p , (14-3) 

n = e »0 n «=0 n *»2 

as is seen by writing out in full. The limits (2 ,«d) on the latter summation 
may be changed to (0,«>) if we agree to define 

= 0 for all negative n. (14-4) 

Hence 

x 2 y" = 2a n (n + p)(n + p - l)x n4 " p 
£ 2 /' = Xa n (n -f p)x n+p 
x 2 y = Sa n _ 2 ^ n4 * p 
—p 2 2/ * S — p 2 a n x n4p . 


According to Eq. (14-1), which we wish to solve, the sum of the four terms 
on the left of the above equations is zero. Hence, the same is true for the 
series on the right. Equating to zero the coefficient of x n+p in the sum of 
these series gives 

a n (n + p)(n + p — 1) + a„(n + p) + a n _ 2 — p 2 a* * 0 


1 The novelty is that we allow p to be any number, whereas if (14-2) is an ordinary 
power series, p must be an integer. Since p may be increased at will, the assumption 
oo pt 0 involves no loss of generality. 



161 


SBC. 14] POWER SERIES AND DIFFERENTIAL EQUATIONS 

or, after simplification, 

On[(n -f- p) 2 — p 2 ] + a n _ 2 = 0. (14-6) 

Equation (14-5) is valid for all n. For negative n Eq. (14-5) holds auto- 
matically by virtue of (14-4). The first nontrivial case of (14-5) is called 
the indidal equation; it is obtained in the present example by putting n * 0 
and takes the form 

ao(p 2 — p 2 ) + 0 « 0, indicial equation, (14-6) 

This shows that p = p or p = — p, since a 0 5^ 0. The other values of a n 
are determined from (14-5) in the form 


On 


1 

(n + p) 2 


2* 


04-7) 


The choice n « 1 gives 1 a\ = 0, and hence a n - 0 for all odd n. Also 


a 2 ** 


a A 


Oq 


(2 + p ) 2 
a 2 


&0 


(4 + p) 2 - p 2 [(4 + p) 2 - p 2 ][(2 + p) 2 - p 2 ] 

and so on. In this way it is easily verified that the series corresponding to 
P = P is 


V * aoX 1 ' 


1 


+ 


2(2p + 2) 2-4(2p + 2)(2p + 4) J 

and that the series for p = — p is the same, with ~p in place of p. 

When p is a nonnegative integer, the expression may be simplified by use 
of factorials, as follows. We take a factor 2 from each term of the denom- 
inator and place it with the x in the numerator, obtaining 


y = a 0 x p 1 


(x/2) 2 


+ 


(x/2) 4 


l(p+l) l-2(p+ l)(p + 2) 

If the denominators are now multiplied by p\, there results 
f 1 (r/2) 2 (x/2) 4 

y aoP ' x [pi i-( p+ l)! + 2Kp + 2)l 
and since x p - 2 p (x/2) p , this yields y = a 0 p\2 p J p (x), where 


(14-8) 


w - Z 


(-\) n (x/2) Sn+p 


n-o n!(p + n)l 
1 Provided the denominator in (14-7) does not vanish. See Prob. 7. 


(14-9) 



INFINITE SERIES 


162 


[chap. 2 


The function J p (x) is called the Bessel function of order p. The graphs of 
Jo(x) and J\(x) are shown in Fig. 7. 

Now, the differential equation (14-1) is meaningful even when p is not a 
positive integer, and the series solution (before introduction of factorials) 
is also well defined for general p. It is natural to inquire if we can define p! 
in such a way that (14-9) is meaningful and satisfies (14-1) for p unre- 
stricted. A glance at (14-8) shows that such an extension may not be pos- 
sible when p is a negative integer, but there appears to be no difficulty 
otherwise. 



To obtain an appropriate definition of p\ for arbitrary p, we introduce 
the so-called gamma f unction V(p) defined by 

F(p + 1) = / t p e-*dt, p > 0. (14-10) 

Jo 

This function was discovered by the celebrated Swiss mathematician L 
Euler. Because of its connection with />’ the function F(p -f- 1) is often 
called th a factorial function and is written p ! or ll(p). We shall use the 
notation pi as soon as we have established that (14-10) gives the appro- 
priate value when p is an integer. For comparison with other notations, 

p! ss Il(p) =* V(p ~f 1). 

If p > t, integration by parts 1 in (14-10) gives 

lOO |00 

I = -fi V I dt. 

Jo 10 Jo 

Since the integrated term drops out, the foregoing relation simplifies to 

r (p + l) » pT(p) (14-11) 

when we use (14-10). 

Writing (14-11) in the form 

Tip) ^p-’lXp-M) (14-12) 

1 Since the improper integrals (Sec. 3) are convergent, the process is justified by writ- 
ing b in place of <*>, carrying out the partial integration, and then letting b -♦ ». Actu- 
ally (14-10) holds for p > — 1, and the partial integration is valid for p > 0. 




POWER SERIES AND DIFFERENTIAL EQUATIONS 


163 


sec. 14] 

enables us to define F(p) for negative values 1 of p. Thus, if any number p between 
0 and 1 is used on the left side of (14-12) then the right side gives the value of F(p), 
for when p > 0, F(p + 1) is determined by (14-10). If the recursion formula (14-12) is 
used again, the values for — 1 < p < 0 can be found from those for 0 < p < 1. That 
is, p -f* 1 in (14-12) ranges over the interval (0,1) if p ranges over the interval ( — 1,0). 
Similarly, when we know F(p) for —1 < p < 0, we can find F(p) for —2 < p < —1, 
and so on. 

Inasmuch as (14-10) gives 

r(l) * | i t Q e,~ t dt « -e-'j - 1, (14-13) 

Jo lo 

the method fails for p*0. Thus 

lim F(p) « lim p“ l F(p -f 1) «- -f « or — °o 

V - 0 P -*■ 0 

according as p — ► 0 through positive values or through negative values. Similar be- 
havior is found for all negative integers, and hence the 
graph of r(p) has the appearance shown in Fig. 8. 

However, by use of (14-10) and (14-1 i) it is easily 
verified that F(p) never vanishes, and hence, if we 
agree that 1/F(p) « 0 for p a negative integer, it wall 
follow that the function 1/F(p) is well behaved for 
every value of p without exception 2 

Equations (14-11) and (14-13) give F(2) » T(l) ~ 1 
and, in succession, 

r(3) - 2 r( 2 ) * 2-1 
r(4) * 3l’(3) - 3-2 1 

T(n + 1) » nl\n) - u(u - 1) ... 3-2-1. 

Hence, the definition 

pf * T(p + 1) (14-14) 

furnishes the desired generalization; it gives a meaning to p ? for all p except p ** — 1, 
—2, —3, . . . , and it gives the familiar value when p is a positive integer. The properties 

p! *= p(p — 1)! or p ~ « (p 'J . " jjl (14-15) 

are ensured by (14-11). The former fails when p is zero or a negative integer, but the 
latter holds for all p, without exception 

The result of this discussion is that 1/p! is well defined everywhere, and 
p! is well defined except for p — —1, —2, —3, . . at which values 1/p! 
vanishes. Moreover, we have the fundamental formula (14-15). When a 
series containing factorials is obtained by solving a differential equation, it 
is almost always this relation (14-15) that makes the series a valid solution, 

1 Equation (14-10) does not serve this purpose, because the behavior of t p at l m 0 
makes the integral diverge when p < —1. 

* The relation of F(p) and l/r(p) is quite analogous to the relation of esc p and sin p, 
respectively. 




INFI mr% SERIES 


164 


[chap. 2 


and hence the extended definition of pi may be used without hesitation. 
In particular, J P (x) and J-. P (z) are solutions of (14-1) no matter what 
value p may have. 

For most values of p the functions J p {x) and J~ p (x) are independent, by 
Theorem I, Sec. 13, and the general solution of Bessel’s equation is 


y = C\J P (x) + C 2 J- p (x ), Ci and c 2 const. 

If p « 0, however, then the two roots of the indicial equation are both 
p 0, so that we obtain only the single function Jo(x). Another excep- 
tional case arises when p = ±1, db2, .... Although the series (14-8) is 
meaningless when p is a negative integer, the series (14-9) is well defined 
and satisfies 


J~m(x) - (-1 ) m J n (x), m ~ 0, 1, 2, 3, 

To see this, we observe that 

A ( — l) w (r/2) 2w ~~ yw = • (— l ) n (x/2) 2w ~~ m 
^ X n^To n\(n-m)\ ", » !(n - m) ! 


(14-16) 


(14-17) 


since the factor l/(n — m) \ is zero when n < m. If the sums (14-9) and 
(14-17) are written in full, then (14-16) follows at once. 1 Because of (14-16) 
the functions J r „ m (x) and J m (x) are dependent, so that we obtain only one 
solution rather than two. 

This failure of the method to provide both solutions is not a serious 
shortcoming, since the second solution can always be found from the first 
by the method of Chap. 1, Sec. 29. Carrying out the calculation in the 
general ease yields the following theorem: 2 

Theorem IT. When the root? p { and p 2 of the indicial equation are distinct 
and do not differ by an integer , the nuthod of Theorem 1 yields two linearly 
independent solutions » If the roots do differ by an integer a second solution 
can be found, by assuming that 

CO 

2/2 = cyi(x) log t +■ £ a n x n+flt (14-18) 

»-*() 


where yi{x) is the solution given by Theorem I for the root p = p\. 


By setting i/i(.r) « 
equation for p = 0 is 

Kq(x) * 


«A»lr), for example, one can show that a second solution of Bessel’s 

(-DV/2) 2 * . 


* Jo(x) log*-S 


{W 


(■>-() 


1 It is suggested that the reader verify this statement by actually writing the sums 
in full 

* It was shown by Frobenius that the second solution can also be obtained by differen- 
tiating the first solution with respect to the exponent p; cf. Piaggio’s book cited in 
Sec* 13. 



SBC. 14] POWER SERIES AND DIFFERENTIAL EQUATIONS 165 

This function is called the Bessel function of the zeroth order and second kind. Thus, 
the general solution of 

xy " -f- / 4 xy * 0 

is y m Cj/ 0(2) 4 c%Ko(x). Other functions K m (x) of the second kind are obtained simi- 
larly. By considering linear combinations of J tn (x) and K m {x) we get the modified Bessel 
functions of the first and second kinds, denoted in the literature by I m (x ), Y m (x), N m (x), 
and H m (x), 


PROBLEMS 

1. Show that Jq(x) * —J\{x) and also that 

— X n Jn(x) * zVn- l(r), ~ X~V*(:r) ** ~ x ~*Jn+ 1&). 

Deduce that 

- Jn+xU) - 2J&r), 

2 n 

J n-l(r) 4 J n 4 l(x) «* — J „(x). 

3 * 

2 . The confluent hyper geometric equation is 

xy" 4 p/ « xy' 4 qy, 

where p and 9 are constant. 

(«) According to Theorem I, what range of validity do you expect for a solution of the 
form 2a»x n+p ? 

(6) Assuming that 1/ ~ ' 2 a n x n '* p , verify that xy" * 2 a„_n(n 4 p 4 l)(n 4 pk’^and 
find similar expressions for py', ry\ and qy 

(c) By considering the coefficient of j n+/> , deduce 

On m( w 4 p 4 1 )(n 4 p 4 p) 3 o n (n 4 p 4 ?) 

for all values of n. 

3 . In Prob, 2, (a) show that the roots of the indicial equation are p « 0 and p ** 1 — p. 
Hint 4 The first nontrivial case arises when n =* - 1. 

(6) When p =* 0, show that 

n 4 q 

a„4-i a „ ^ j)( n p ) 

if p is not zero or a negative integer. Thus get the solution 

r . o (g + n J* g(? + 1)(<? + 2) i 3 -j 

Vl - 00 p + p j, + - (p + , )2 , + p( p '+ i)( p + 2) 3' + ” J 

(c) Similarly, obtain the solution corresponding to p * 1 — p when p is not a positive 
integer. 

4 . For the hypergeo metric equation of Gauss, 

r(l ~ x)y" 4 [e — (a 4 & 4 Ik]?/ - ahy « 0, 
obtain one solution in the form 

* r(o 4 n) r (b 4 n) 
y ntTo r(l 4 n) V(c 4 n) 


when a and b are not negative integers. 



166 

$. For the equation 


INFINITE SERIES 


[CHAP. 2 


xy" + (1 - 2 p)y' +xy ~ 0 


obtain a recursion formula for the coefficients of a series solution, and show that one 
solution is y » x p J p (x). 

6* As in Prob. 6, show that y » x h J p (\x) satisfies 

4 x 2 y" + (4X 2 jc 2 — 4 p 2 -f l)y = 0. 

7* (a) Given that r(k£) «* V*, obtain the formulas 


Jm(x) 






(b) What is the general solution of Bessel's equation with p ** H? (This shows that 
Theorem I may yield the general solution even if w ~~ pi is an integer.) 

8 . The generating function of the sequence J n (x) is 

e 2 ^ ~ h) „ £ J n (x)h n . 

n=>— oo 

Verify that the coefficient of h° in the expansion of the exponential is, in fact, ./oO). 
Hint: By the series for e u the exponential is 

=©‘(‘- 0 ‘=- 

Pick out the term independent of h in the binomial expansion of [A — (1/^)]'* when n 
is even, and note that there is no such term when n is odd. 


SERIES WITH COMPLEX TERMS 

16. Complex Numbers. The equation x 2 + 1 = 0 cannot he solved by 
means of real numbers because the rule of signs does not allow the square 
of a real number to be negative. But if one adjoins a symbol i to the real 
numbers, which satisfies the equation 

i 2 =~l (15-1) 

by definition, then one can construct the so-called complex numbers a + In 
The latter satisfy the algebraic laws obeyed by real numbers, and they in- 
clude the real numbers as a special case. Moreover, complex numbers en- 
able us to solve not only the equation x 2 + 1 = 0 but every polynomial 
equation. 

Since we want to keep the familiar laws of algebra, it is easy to see how 
addition and multiplication of complex numbers ought to be defined. In- 
deed, 1 if a, b , and i in a + ib are to be treated like any other numbers of 
elementary algebra, then 

(a + ib) + (c + id) = (a + c) + i(b + d). (15-2) 

1 See also Chap. 7, which contains a discussion of complex numbers and functions from 
a somewhat different point of view. 



SEC. 15] SERIES WITH COMPLEX TERMS 167 

This equation is now taken as the definition of addition. In the same way 
we are led to define multiplication by 

(a + ib)(c + id) = ( ac — bd) + i(bc + ad), (15-3) 

since elementary algebra would give Ihe product 

ac + ibc + iad + i 2 bd = ac + i 2 bd + i(bc + ad), 

and (15-1) asserts that i 2 = —1. Finally, we agree that a + ib = a + hi. 

It is easy to verify that these definitions (15-2) and (15-3) do preserve the familiar rules 
of algebra (including those rules that were not considered iu fi anung the definitions) For 
example, complex numbers Zk satisfy 

Z\ 4* 2*2 *** 22 4“ *1, (%l 4* 4“ 23 ~ 2l + (?2 4" Zj), 

Z1Z2 * 22^1, (ZlZ^Z'J « 21(2223), 

21(22 *F 23) = 2|2'2 4 - 2iZ 3 . 

Also there is a zero, and there is a unit: 

z 4 (0 4 - Oi) ** 2, 2(1 4 - Or) «= z for all z. 

Moreover, the complex number a 4- 0i is found to be equivalent in every respect, except 
notation, to the real number a Hence in this sense the complex numbers contain the 
reals as a special cast*, and we have a right to consider that 

a ~b Oi - a. (15-1) 

The convention (15-4) also agrees with our purpose of Keeping the rules of algebra intact. 
Using (15-4) we write 0 and 1 for the zero and unit element of our algebra. Subtraction 
is defined by considering th»* equation (a 4- ib) -f 2 - 0; it will be found that 

-(a-f bi) * (~a) -4 (—b)i ** (~l)(a 4 - tb). 

Although we made no attempt to preserve the euricelhition law, it is nevertheless valid; 
that is, 

Z 1 Z 2 = 0 only if zi =* 0 01 22 ** 0. 

And finally, the possibility of division is suggested by 

a 4- ib (a 4 tb)(c — id) 
c 4 id (r 4- id)(c — id) 


ac 4- bd be — ad 

= CHM* + ’ c 2 T?' 

Now, the latter expression can be shown by (15-3) to satisfy the equation 

(c 4- id)z « a 4- ib. 

Hence, the result of this heuristic calculation is, in fact, the quotient. Thu process breaks 
down if c 4* id «* 0, but only then. 

The general tenor of this discussion is that the algebra of complex num- 
bers agrees with the algebra of real numbers and we need not hesitate to 



168 


INTINITE SERIES 


[CHAP. 2 

apply the familiar rules to the new symbols. There is one new feature, 
however. When we say that a ib is a symbol for a complex number, we 
assume, naturally, that different symbols represent different numbers. In 
other words, if 

a + ib * a' + ib' for a, a', b , b' real, then a * a' and b *= b'. (15-5) 

This important relation may be taken as the definition of equality. Un- 
like the algebraic properties described hitherto, (15-5) is true for complex 
numbers only; it does not hold if i is replaced by a real number. 

The following alternative analysis of (15-5) shows the role of the equation i 2 «* — 1 
and also shows why a, a b , b' must be assumed real. If a -f ib «* a' -f ib\ then 

a ~~ a' — i(b f — b). 

Squaring gives (a — o') 2 ** —(6 — b') 2 because i 2 « —1, and hence 

(a - a') 2 + (b - b') 2 - 0. 

Since the square of a real number is positive unless the number is zero, the latter equa- 
tion implies a — a' ** 0 and b — b' = 0. 

Complex numbers z = x + iy may be represented graphically in the so- 
called z plane by introducing two perpendicular axes, one for x and one for 
y (Fig. 9). The x axis is called the real axis , and x is the real part of x + iy; 
the iy axis is the imaginary axis, and iy is the imaginary part of x -f iy. 
The absolute value of z is the distance from the representative point to the 
origin; it is denoted by \z\, as in the case of real numbers. Evidently, the 
points satisfying \z\ = r lie on a circle of radius r centered at the origin. 
The interior of this circle consists of the points \z\ < r. When z = x + iy, 
then 

|*| - V? + 'y*. (15-6) 

A short calculation based on (15-6) gives 

Z\ \Z\\ 

I *1*2 1 = |*1 1 |*2l, — =7—7 if*2^0, 

z 2 i z 2 I 

so that the absolute value of a product is the product of the absolute values, 
and similarly for quotients. 

Since real and imaginary parts are added separately, computation of 
zt + z 2 can be effected as shown 1 in Fig. 9. Inasmuch as the sum of two 
sides of any triangle is greater than the third, the figure gives the important 
inequality 

|*i + **l < l*il + l**|. (15-7) 

1 One may think of z * x + iy as a vector with components x and y. The method of 
adding vectors by adding components agrees with the definition (15-2), and hence the 
construction of Fig. 9 is simply the parallelogram rule familiar from mechanics. 



SRC, 16} SERIES WITH COMPLEX TERMS 169 

A similar result may be obtained from this one for any number of complex 
numbers. 

If s n * u n + iv n and s « u + iv, we define lim s n =» s to mean that 
simultaneously 

lim u n « u, lira v n *= v. (15-8) 



This shows that the theory of limits for complex numbers can be based on 
the corresponding theory for real numbers. 


PROBLEMS 


1. Show that | Z 1 Z 2 i « jzi | 1 22 !. 

2. Show that hm s n * & if, and only if, lim — aj =0. 

3. Sketch the set of points z in the complex plane described by (a) | z | « 1, (b) \z\ <2, 
(r) \ z \ > 1, (d) 1 2 — 2 1 <1. Hint: j z — a | is the distance from the representative 
point for z to that for a. 

16* Complex Series. Convergence of infinite series of complex numbers 
is defined by considering the limit of partial sums, just as for real series. 
By (15-8) the complex series converges if, and only if, the two real series 
obtained by considering real and imaginary parts are both convergent. 
In other words, 

X(p n + iq n ) - V + n 

if, and only if, Xp n = p and X q H - </. Because of this correspondence of 
real and complex series, most of the results presented hitherto in this chap- 
ter apply with little change to the complex case, and the proof also involves 
nothing new. 

As an illustration let us show that the general term of a convergent senes approaches 
zero . If a n is complex, we have 

n n~~ 1 

a n ** £ ~ 23 a k> 

ft- 1 *-l 

and taking the limit as n -> « yields lim a n » 0, exactly as in the proof for real series. 
Alternatively, one may use the result for real series Namely, if a* » p* -f then the 



INFINITE SEMES 


170 


[chap, 2 


convergence of 2a* implies the convergence of 2p* and of 2<?*. Hence, by the theorem for 
real series, lim p n « 0 and lim g„ ■» 0. Consequently, lim a n — 0. 

As a second illustration, we show that an absolutely convergent senes is convergent. 
With a* ** pa + iqic, we have 

|P*I “ Vpj < Vp| + r/l - |o*|. 


Hence, if 2 1 a* | converges, then 2 1 p* | converges by the comparison test for real series, 
and then 2p* converges, because we know that for real series absolute convergence im- 
plies convergence. Similarly, 2 <ik converges, and hence 2 (pa ~r ag) converges. 

As a third example, the leader may obtain the analogue of Theorem I, See S, foi com- 
plex series. That is, if 2 a n z n converges for z — zo, then the seiies converges absolutely 
for all z such that \z\ < (zo I and uniformly for all z such that \z\ < \z\\ < |*n| It will 
be found that the proof is the same, word for w ord, as the proof in the ease of real senes 
The symbol for absolute value, however, has the meaning assigned in (15-6). 


For many series 2 u n (z), the set of points z at which the series converges 
gives a complicated region in the z plane. It is a remarkable fact that for 
a power series the region of convergence is always a circle centered at the 
origin. The circle is called the circle of convergence, and the radius of the 
circle is the radius of convergence. We agree to take the radius as zero if 
the series converges for z — 0 only and as infinity if the series converge* 
for all z. At points on the boundary of the circle the series may oil her 
converge or diverge, just as in the ease of the interval of convergence for 
real series. 

For proof that the region is a circle, let 2a n z n be a power series which 
converges for some value z = Zo ^ 0 but diverges for some other value 
z «= z\. As we have already noted, the fact that the series converges for 
z = z 0 makes the series converge throughout the circle \z\ < j z {) j On 
the other hand, the series obviously does not converge throughout, any 
circle containing the point z Y (see Fig 10). We let C be flic largest circle | z | 
= r such that the series converges at every interior point of C. The radius 
r of C is at least equal to \z 0 \ but does not exceed \z x \. 

To show that C is the circle of convergence, all we have to do is establish 
that the series diverges at every point exterior to C. Let z 2 be an exterior 
point, so that \z 2 \ > r. If the series converges at z 2f then it would have to 
converge throughout the circle \z\ < \z 2 \. But this contradicts the fact 
that C is the largest circle throughout which the series converges, and hence 
the proof is complete. 1 

To illustrate the concept of circle of convergence consider the series 

— — 5 = 1 - x 2 + x* H h (-l)V n H 

1 + x l 

1 The fact that a largest circle exists is quite clear from the geometry. An analytical 

proof may be given, if desired, by constructing circles with successively larger radii 
and using the fundamental principle (Sec. 1). 



SEC. 17] SERIES WITH COMPLEX TERMS 171 

which converges for \x\ <1 but diverges at x = dbl. If 1/(1 + x 2 ) is 
regarded as a function of the real variable x , there appears to be no reason 
why the series should diverge when \x\ >1, for 1/(1 -f x 2 ) has derivatives 
of all orders at every value of x. But when we regard x as a complex vari- 
able, the divergence is explained by the fact that the denominator l + x 2 
vanishes at x = dri. Clearly, if the circle of convergence cannot contain 
the points dti, then the radius of convergence cannot exceed 1. 



Fig 10 


PROBLEMS 

00 

1. Verify that the series £ z n frv eon verges absolutely at every boundary point of 

1 

00 

its circle of convergence whereas 7i 2 z n converges at no boundary point of its circle 

1 

of convergence 

2. If /(*) - and /( 0) - 0, 

(а) How does f(z) behave when z «= x and x 0 through real values? 

(б) How does f{z) behave when z * ly and t/ — ► 0 through real values? 

(c) Could this function have a power-series expansion valid m some circle contain- 
ing the origin? Explain. 

17. Applications. By means of power series the functions sin e x , log x, 
tan^ 1 x, J p (x), and so on, may be extended to a complex variable z. For 
example, since the series 

/j*2 £.3 

— log (1 ~ x) x + “• + —- H 1 1 

2 3 n 



172 INFINITE SERIES [CHAP. 2 

converges for Jx| < 1, we know without further discussion that 

z 2 z 3 z n 

z + “ + -*• H 1 1 

2 3 7i 


converges for |zj <1. The latter series is the definition of — log (1 — z). 
Similarly e* y sin z t and cos z are defined by 


e m » 



sin z = 


(2n+ l)T 


cos z = 


2 


(~I)V W 

1^)7“ 


Many familiar formulas can be extended at once to the complex-variable 
case. To establish that 


sin 2 z + cos 2 z ~ l t 


(17-1) 


for example, we reason as follows. It is known that (17-1) holds when z is a 
real variable x, and hence 


(-l) n x 2n+1 1 

2 

, (-I)"* 2 "] 

2- 

+ 

V x 

(2 n -f* 1) ! _ 

L (2 n) ! J 


(17-2) 


when x is real. The left side of (17-2) is a power series, as we see by imagin- 
ing that the terms are collected. Since the power series is zero for an inter- 
val of x values, every coefficient must be zero (Sec. 8, Theorem III). Hence 
the power series is also zero when x is replaced by a complex variable z, and 
this establishes (17-1). 


The same method may be used to prove formulas involving two variables; for example, 
from 

e xi « e*'e r * for X\ and x% real, 
it follows without detailed calculation that 

e*i+*2 «, e *i e *2 f or Zl am j Zj complex. 

The systematic development of this idea leads to a branch of analysis known as the 
theory of analytic continuation. 


Upon letting z * ix in the series for e g , we get 

(• ix) n - 2 ~ 4 ~ 6 ' ~ 3 


x .„, x xr x w 

sss 2J ass 1 — 1~ — — f~ * 

nl 2! 4! 01 


.(* 


when we write out the sum in full, noting that 




■h 


♦ * -t, 


1, 


x x 
1 

3! 51 


n =*= t f 


7 ! 


+ * 



6UC. 17] SERIES WITH COMPLEX TERMS 173 

The series representations for cos x and sin x now give Euler’s formula, 

e % * = cos x + i sin x , (17-3) 

which expresses the exponential function in terms of the trigonometric 
functions. On the other hand (17-3) also leads to 

e ix + e -ix e ix _ 

cos x = , s j n x « : , (17-4) 


as the reader can verify. These equations are constantly used in the study 
of periodic phenomena, for example, in network analysis and synthesis, in 
physical optics, and in electromagnetic theory. The calculations are ordi- 
narily carried out by complex exponentials, and then the appropriate 
trigonometric form is obtained by taking real or imaginary parts [cf. (17-3)]. 


Example 1. Obtain the trigonometric identity 


cos u -f cos 2u H h cos nv = 


sin ( n 4- 4£)« 
2 sin Yu 

for u 9 * 0, ±2*-, ”fc4ir, . . . , by considering the exponential sum 
8 - <? 1U + e 2lu + f e n ' u . 

The series 8 is a geometric series with ratio r * c*“, and hence 


1 


(17-5) 


(17-6) 


by (1-18). If the numerator and denominator in (17-6) are multiplied by e we get 

ptnu _ ! ^ u — e t ^ u 

_ e ~thv 2? sin l /iu 

upon using (17-1). By (17-3) this yields another expression for the exponential sum s, 

cos (n 4" Y)a 4* f sin (n -f Mi) a — cos \<iu — x sin Yu 
2i sin Y^u 

which leads to (17-5) when wo equate real parts. 

Example 2. Show that / e %kx dx « 0 if k is a nonzero integer, and deduce 


i: 

/: 


• cos mx dx 


cos nx sin mx dx *= 0 


[** . . . JO, 

J sin nx sin mxdz » 


if m n, 
if m »* n, 


if m n, 
if m *» n, 


(17-7) 


whenever m and n are positive integers. 



[chap. 2 


174 


INFINITE SERIES 


If k is a nonxero integer, (17-3) gives 
r n f 2v 


J r* t 2 

F e 4 ** dx ** I (cos kx + i sin kx) dx * 0, (17-8) 

o J o 


which is the first result. By (17-4), 


r** 
i c 
Jo 


•A 


4 / cos nx cos mz dx » / (e tnx -j- c + <T~ tm *) dx 

j j- g *(»«4n)* c »(m— n)« _j_ e t(»-m)a: _|_ € ~ t(*»-f«)xj fa 

Jo 


Each term is of the form e^ 1 with k an integer, and hence we get zero unless m * n. 
If m =* n, the two middle terms give 2, so that the integral is 4ir. The other relations 
(17-7) are established similarly. 


PROBLEMS 


1. By using the series definition for e z , show that ( d/dx)e cr « ce cx when c is a complex 
constant. (Since we have not defined the derivative with respect to a complex variable, 
assume that x is real.) 

2. Sum the series sin x + sin 2x H — ■ + sin nx. 


. Evaluate j e ax cos bx dx and j e ax sin bx dx by considering yV (o+t6)l dx and equating 

e (a+ib)z € <** e *x( a _ ib ) 


real and imaginary parts. Hint: 

,2T a + lb 


r ■ ° 2 + 62 
4. Evaluate / (2 cos x) 4 dx by using the formula 

Jo 


2 cos x =* e xx -}■ 

together with (17-8). 

6. Show* that every complex number z may be written in the form z *» re 1 *, where 
r > 0 and 0 < 9 < 2r. Bird : If z « x + ty, introduce polar coordinates ( r,d ), so that 
x » r cos 0 and y =» r sin 9. 

6. (o) For 0 < r < 1 obtain the expansion 


-i *= Sr n (cos n9 + i sin nd) 

J — re* 0 


by letting z *> re* in the series for 1/(1 — z). 

(6) Separate 1/(1 — re*) into real and imaginary parts, by noting that 


1 — re ^ 1 1 — r cos 9 ir sin 0 

1 — re~~ * 1 — re* 1—2 r cos 0 4* r* 


(c) From (a) and (6) deduce that 


1 — r cos 9 
1 — 2r cos 9 -{- r* 
r sin 9 

1 — 2r cos 9 -h r* 


* 53 r " cos 

n»0 

oo 

« 53 r n sin ntf, 

n =— 0 


0 < r < 1, 
0 < r < 1. 


[The first series of (c) is an example of a Fourier cosine series , and the second is a Fourier 
sme series. The study of such series by real-variable methods forms the topic of the next 
eight sections.] 



SBC. 18] 


FOURIER SERIES 


175 


FOURIER SERIES 

18. The Euler-Fourier Formulas. Trigonometric series of the form 

00 

/« ~ o + 23 ( a n cos nx -f & n sin nx) 


n«l 


(18-1) 


are required in the treatment of many physical problems, for example, in 
the theory of sound, heat conduction, electromagnetic waves, electric cir- 
cuits, and mechanical vibrations. An important advantage of the series 
(18-1) is that it can represent discontinuous functions, whereas a Taylor 
series represents only functions that have derivatives of all orders. 

We take the point of view that /(x) in (18-1 ) is known on ( — tt, it) and 
that the coefficients a n and b n are to be found. In order to determine a 0 , 
we integrate (18-1) term by term from — x to tt. Since 


/: 


cos nx dx 

r 

the calculation yields 


s f sin nx dx — 0 for n = 1,2, . . 

j — * 

J f(x) dx * a 0 T. 


(18-2) 


The coefficient a n is determined similarly. Thus, if we multiply (18-1) by 
cos nx, there results 

/(x) cos nx » }4ao cos nx 4 h a n cos 2 nx H , (18-3) 

where the terms not written involve products of the form sin mx cos nx or 
of the form cos mx cos nx with m 9 ^ n . It is easily verified 1 that for inte- 
gral values of m and n, 

f sin mx cos nx dx — 0, in general, 

— T 

f eos rnx cos nx dx = 0, when m ^ dbn, 

J —r 

and hence integration of (18-3) yields 

f f(x) cos nx dx = o n f cos 2 nx dx ** a n r. 


Therefore, 


1 [* 

a n * - / /(x) cos nx dx. 

-tr J —IT 


(18-4) 


By (18-2), this result is also valid ! for n - 0. 

1 See Example 2 of Sec. 17. 

9 That is the reason for writing the constant term as * 4 oq rather than oq . 



170 


INFINITE SERIES 


[chap. 2 


Similarly, multiplying (18*1) by sin nx and integrating yield 


bn 


1 r* 

- / f{x) sin nx dx . 

t 


(18-4 a) 


The formulas (18-4) are called the Eider-Fourier formula#, and the series 
(18-1) which results when a n and b n are determined by the Euler-Fourier 
formulas is called the Fourier series of f(x). More specifically, a Fourier 
series is a trigonometric series in which the coefficients are given, for some 
absolutely integrable function l f(x), by (18-4). 


The distinction between a convergent trigonometric series and a Fourier series is im- 
portant in the modern development of the subject and is a genuine distinction. For 
instance, it is known that the trigonometric series 


DO 

E 


sin nx 
log (I + n) 


is convergent for every value of x without exception, and yet this series is not a Fourier 
series. In other words, there is no absolutely integrable function f(x) such that 



cos nx dx 


0, 



sin nx dx 


TT 

log (1 + n) 


On the other hand a series may be a Fourier series for some function /(x) and yet diverge. 
Although such functions are not considered in this book, they often arise in practice, 
for example, in the theory of Brownian motion, in problems of filtering and noise, or in 
analysing the ground return to a radar system Even when divergent, the Fourier series 
represents the main features of/(x), ami for this reason Fourier series are an indispensable 
aid in problems of the sort just mentioned. 

Treatises devoted to Fourier series commonly replace the sign of equality in (18-1) 
by ££, or some similar symbol to indicate that, the series on the right is the Fourier 
series of the function on the left. We shall continue to use the equality sign because the 
series obtained in this book do, in fact, converge to the function from which the coeffi- 
cients were derived. 


To illustrate the calculation of a Fourier series, let f(x) ** x. By Eqs. 
(18-1) i 

a n = - j* x cos nx dx ~ 0, 


i r 

6n = -/_ 

TT J 


2 2 

x sin nx dx = cos nir * - ( — l) n+1 , 

n n 


so that, upon substituting in (18-1), 


— 2 ^sin x 


sin 2x sin 3x 

— T“ + — 


(38-5) 


In Sec. 24 it is shown that the series (18-5) does converge to x for — n < 
x < TT. To discuss the convergence outside this interval, we introduce the 
notion of periodicity. A function fix) is said to be periodic if fix + p) 
1 This means that j f(x) J, as well as f(x), is integrable. 



FOURIER SERIES 


SEC, 18] 


177 


* f(x) for all values of x, where p is a nonzero constant. Any number p 
with this property is a period of f(x ) ; for instance, sin x has the periods 2t, 
— 2ir, 47r, .... 


Now, each term of the series (18-5) has period 2 tr, and hence the sum 
also has period 2 ir. The graph of the sum therefore has the appearance 
shown in Fig. 11. Evidently, the sum is equal to x only on the interval 
— ir < x < 7r, and not on the whole interval — oo < x < a>. 

It remains to describe what happens at the points x = zbx, db37r, . . . , 
where the sum of the series exhibits an abrupt jump from — 7 r to -f-ir. 



Upon setting x = dzir, rb37r, ... in (18-5), we see that every term is zero. 
Hence the sum is zero, and this fact is indicated in the figure by placing a 
dot at the points in quest ion. 

The term a n cos nx 4- sin nx in (18-1) is sometimes called the nth harmonic (from 
analogy with the theory of musical instruments). The first four harmonics of the series 
(18-5) are 

2 sin x, — sin 2 j, % sin 3 j, — sin 4x. 

These and the next two harmonics are plotted as the numbered curves in Fig. 12. The 
sum of the first four harmonics is 

y « 2 sin x — sin 2x + % sin 3x — sin 4x. 

Since this is a partial sum of the Fourier series, it may ho expected to approximate the 
function x . The closeness of the approximation is indicated by the upper curves in Fig. 
12, which show this partial sum of four terms together with the sums of six and ten 
terms As the number of terms increases, the approximating curves approach y «■ x for 
each fixed x on — ir < x < 7r but not for x = 

The foregoing example illustrates certain features which are character- 
istic of Fourier series in general and which will now be discussed from a 
general standpoint. Each term of the series (18-1) has period 2ir, and hence 
if /(x) is to be represented by the sum, then/(x) must also have period 2*\ 
Whenever we consider a series such as (18-1), we shall suppose that/(x) is 
given for — ir < x < ir and that outside this interval /(x) is determined by 



178 


INFINITE SERIES 


(CHAP. 2 



the periodicity condition 

f(x + 2ir) = f(x). 

Of course, any interval a < x < a -f 2* would do equally well. 

The term simple discontinuity 1 is used to describe the situation that 
arises when the function f{x) suffers a finite jump at a point x = x 0 (see 
Fig. 13). Analytically, this means that the two limiting values of f(x), as 
x approaches x 0 from the right-hand and the left-hand sides, exist but are 
unequal ; that is, 

lim/(x 0 + e) lim f(x 0 — c), € > 0. 

« — o « —• o 

In order to economize on space, these right-hand and left-hand limits are 
written as /(x 0 +) and /(x 0 — ), respectively, so that the foregoing inequal- 
ity can be written as 

/(* 0 +) 5 * /(*<>-)• 

A function /(x) is said to be bounded if the inequality 

!/(*) I < M 

1 For an example of a discontinuity which is not simple, consider sin (l/x) near x «■ 0. 




SEC. 18 ] 


FOURIER SERIES 


179 




holds for some constant M and for all x under consideration. For example, 
sin x is bounded, but the function 

1 

f(x) « -» for x & 0, 
x 

m - o, 

is not, even though the latter is well defined for every value of x. It can 
be shown that if a bounded function has only a finite number of maxima 
and minima and only a finite number of discontinuities, then all its dis- 
continuities are simple. That is, f(x+) and f(x — ) exist at every value of z. 

The functions illustrated in Figs. 11 and 13 satisfy these conditions in ever; finite 
interval. On the other hand, the function sin(l/jr) has infinitely many maxima near 
x * 0, and as we have noted, the discontinuity at x * 0 is not simple. The function 
defined by 

f(x) = x 2 sin -f x 0, 
x 

/( 0 ) * 0 


also has infinitely many maxima near x *= 0, although it is continuous and differentiable 
for every value of x. The behavior of tnese two functions is illustrated graphically in 
Figs. 14 and 15. 



Fig. 15 


180 


INFINITE SERIES 


[chap. 2 

With these preliminaries, we can state the following theorem, which 
establishes the convergence of Fourier series for a very large class of func- 
tions: 

Dirtchlet’s Theorem. For < x < r suppose f{x) is well defined , is 
bounded , has only a finite number of maxima and, minima , and has only a 
finite number of discontinuities . Let f(x) be defined for other values of x by 
the periodicity condition fix + 2v) » f(x). Then the Fourier series for f(x) 
converges to 

y 2 [f(x+) +/(*-)] 

at every value of x [and hence it converges to f(x) at points where f(x) is con- 
tinuous]. 

The conditions imposed on f(x) are called Dirichlet conditions , after the 
mathematician Dirichlet who discovered the theorem. In Bee. 24 we estab- 
lish the conclusion under slightly more restrictive conditions which are 
sufficient, however, for almost all applications. 

Example 1. Find the Fourier series of the periodic function defined by 
f(x) * 0, if -x < x < 0, 

f(x) =» x, if 0 < x < x. 

By (18-4) we have 

0 dx J x ** x, 

1 f v 

a n *■ — / x cos nx dx ** 0, n > 1 , 

IT JO 

1 f* 1 

b n « - / x sin nx dx » ~ (1 — cos nx). 
x Jo W 

The factor (1 — cos nr) assumes the following values as n increases: 


n = 

1 

2 

3 

4 

i 

5 

... 

(1 — cos nx) » 

2 

0 

2 

0 

2 

. . . 


Determining b n by use of this table, we obtain the required Fourier series 


■4-2 


( sin 

“I 


x sin 3x 
~ H ;r~ 


sin hx \ 


The graph of fix) consists of the x axis from —x to 0 and of the line AB from 0 to x 
(Fig. 16). There is a simple discontinuity at x « 0, at which point the series reduce! 
to */2. Since 

1 IJLr /(°-“) 4/(04), 

2 " 2 


this value agrees with Dirichlet's theorem. Similar behavior is found at x ** ±x, 
sfc2ir, .... 



SBC. 18 ] 


FOURIER SERIES 


181 



The figure shows the first four partial sums, whose equations are 

sin 3x\ 


X X / 

V * » -f 2 sin x, y * 2 + 2 V 


sin x 4* - 


^sin 


. sin 3x sin 5x' 

+ 2 ( sin z -f — — + — r 
3 o 


) 


For most functions it is only the infinite series that reduces to J^j /(x~) ■+•/(* 4*) 1 at 
points of discontinuity. In the present example, however, this condition is satisfied by 
the partial sums, as the reader can verify. That is, the graph of each partial sum contains 
the points (0,x/2), (:i:x,x/2), .... 

Example 2. Find the Fourier series for the periodic function /(x) defined by 
f(x) « -x, if -x < x < 0, 

f(x) « x, if 0 < x < x. 


The integral (18-4) may be expressed as an integration from — x to 0, followed by 
integration from 0 to x. If the appropriate formula for /(x) is used in these two inter- 
vals, we get 


On 


x cos nxdx + J x cos nx dx j 


■:(£ 

1 / cos nx 1 \ 1 / 

; (° + — - nd m Z ( 


cosnx — 1> 


The integration assumes that n ^ 0; if n » 0, we get oq * — r/2, as the reader can 
verify. Similarly, 


~ ** 4“ J * sin no* dx^ 

- - cos nx cos nx^ ® — (1 —2 cos nr) 

r \n n n / n 



182 

Therefore 


INFINITE SERIES 


(CHAP. 2 


fix) 


2 cos 3x 2 cos bx 


n . sin 2a: 3 sin 3a? 

+ 3 sin x + — 


sin 4a; f 3 sin 6a; 

A * f- 


When x » 0, the series reduces to 

x 2 /J_ 

4 


i i 

+ p +^+- 


which must coincide with (soe Fig. 17) 

/( 0 +)+/( 0 -) 

css — • 

2 2 



Thus 

Hence 


! / 1 1 1 \ 

Vl J + 3 J + 5 2+ '/ : 


111 jr 2 

P + P + 5 S+-" " g- 


This example suggests the use of Fourier series in evaluating sums of series of constant 


PROBLEMS 

1. Evaluate J cos mx cos nx dx for integral m and n by use of the identity 

2 cos A cos B » cos {A -f B) -f cos ( A — B). 

2. Find the Fourier-series expansion for f(x) } if 


m - =. 

for — r < x < 

2 

m - o. 

, r 

for - < x < w. 

m - -x. 

for — ir < x < 0, 

f(x) - 0, 

for 0 < x <vr t 




6. In Probfl. 2 to 5, sketch the graph of the function to which the Fourier series con- 
verges in the range — 4ir < x < 4v, 

19* Even and Odd Functions. For many functions the Fourier sine or 
cosine coefficients can he determined by inspection, and this possibility is 
now to be investigated. A fund ion f(x) is said to be even if 

f(-x) s /(*), (19-1) 

and the function fix) is odd if 

/(-*) - —/(*)• ( 19 - 2 ) 

For example, x 2 and cos x are even, whereas x and sin x are odd. The 
graph of an even function is symmetric about the y axis, as shown in Fig. 
18, and the graph of an odd function is skew symmetric (Fig. 19). By 


inspection of the figures it is evident that 



f f(x) dr = 2 f f(r) dr 

if f{x) is even, 

(19-3) 

f f{x) dx — 0 if /(x) is 

J —a 

odd, 

(19-4) 



Flo. 18 



184 


INFINITE SERIES 


[CHAP* 2 



Fia. 19 


since the integrals represent the signed areas under the curves. 1 For 
example, 

f sin nx dx = 0, 

J —a 

since sin nx is an odd function. 

Products of even and odd functions obey the rules 

(even) (even) = even, (even) (odd) = odd, (odd) (odd) = even, 

which correspond to the familiar rules 

(+1X+D « +1, (+i)(-D - -l, (-D(-l) - +L 

For proof, let F{x) — f(x)g(x) y where /(x) and g(x) are even. Then 

F(-x) = f(~x)g(~x) * f(x)g(x) * F(x) f 

which shows that the product f{x)g(x) is even. The other two relations 
are verified similarly. As an example, the product cos nx sin mx is odd 
because cos nx is even and sin mx is odd. Hence, (19-4) gives 

f cos mx sin nx dx «* 0 
J —a 

without detailed calculation. 

The application of these results is facilitated by the following theorem : 
Theorem. If /(x) defined in the interval — v < x < ir is even , the Fourier 
series has cosine terms only and the coefficients are given by 

2 /■*■ 

a n * - / /(x) cos nx dx y b n » 0. (19-5) 

t y o 

1 An analyse proof of (19-3) and (19-4) may be based on (19-1) and (19-2). 





INFINITE 8EEIES 


186 


[chap. 2 


The function to which this series converges is illustrated in Fig. 22, and the sum of the 
series (19-7) is presented graphically in Fig. 11. 



Since \x\ — x for x > 0, the two aeries (19-7) and (19-8) converge to the 
same function x when 0 < x < tt. The first expression (19-7) is (‘ailed the 
Fourier sine series for x , and (19-8) is the Fourier cosine series. Any func- 
tion fix) defined in (0,ir) which satisfies the Dirichlet conditions can be 
expanded in a sine series and in a cosine series on 0 < x < r. To obtain a 
sine series, we extend fix) over the interval — it < x < 0 in such a way that 
the extended function is odd. That is, we define 

Fix) = fix) on 0 < x < r, 


Fix) = — /( | x | ) on — 7r < x < 0. 

The Fourier series for Fix) consists of sine terms only, since Fix) is odd. 
And the coefficients are given by (19-6) because Fix) = fix) on the interval 
0 < x < 7r. Similarly, if it is desired to obtain a cosine series for fix) on 
0 < x < the coefficients are given by (19-5). 


Example' Obtain a cosine series and also a sine series for sm x. 

For the cosine series (19-5) gives b n » 0 and, after a shoit calculation, 

2 [ T . 2(1 -f cos vie) 

a„ » - I smxcosnxax * — » n ^ 1. 

tt Jo ir(l — n z ) 

For n ** 1 the result of the integration is zero, and hence 


2 4 / cos 2 x cos 4 x cos Ox \ 

sm x I — + b I 

TT \2 2 ~ 1 4 2 - 1 6 2 - 1 / 


when 0 < x < ir. Since the Rum of the series is an even function, it converges to | sin x | 
rather than sinx when — t < x < 0. This shows, by periodicity, that the series con- 
verges to | sin x | for all values of x. 

To obtain a sine series, (19-6) gives a n ** 0 and 


, 2 r. . . jo, 

b n ** - / sm x sm nx ax ** < 

T Jo ll» 


for n > 2, 
for n » 1. 


Hence the Fourier sine series for sin x is sin x, just as one would expect. That this is 
not a coincidence is shown by a Uniqueness Theorem: If two trigonometric series of 


FOURIER SERIES 


SEC. 20] 


187 


the form (18-1) converge to the tame eum for all values of x, then corresponding coefficients 
are equal} 


PROBLEMS 


1. Classify the following functions as even, odd, or neither: 

1 4- x 

x 2 , x sin x, x * cos nx t x A % log » e*, g(x % ). 

1 — x 

2. Prove that any function can be represented as the sum of an even function and an 
odd function. Hint: f(x) = l A[f(x) -f /(-*)] + t 2 [/(x) ~ 

3. For 0 < x < * show that 


x sm 3 j sin 5 x 

_« 81ni + __ + ___ i . 

Hint ‘ Take/(j) « *74 in (19-6) 

4. A function is defined by /(x) = xforO < x < x/2and//» = 0 elsewhere in (—x,x). 
Find the Fourier series, the Fouriei sme series, and the Fourier cosine series. In each 
case sketch the graph of the sum of the series lor —4 it < x < 4tt. 

6. By taking /(x) — x 2 in (19-5^ show that 


+ 4l(-ir ! 


for — x < x < x, and deduce that 

x 2 1 
12 ^ 


l jl _ 1 

i 2 ~ 2* + 3 2 4 2 + ’ 


6. Show that if f(x) ~ x for 0 < x < x/2 and f(x) = x — x for x/2 < x < x, then 


/W 


( cos 

7 


cos 2x cos fix cos 1 Ox 

+ — 2 - + --^-+- 


7. Show that for — x < x < x, 

sin ira 


COS Otx 

when a is not an integer. Deduce 


HU -A 2 a sin x« 

— + Z “I) .- r — ,-cosnx, 

* i ir(a 2 - n 7 ) 


1 (\ * 2a \ 

cot X« *= — I - — '•> 9 I * 

x \a n*i W 2 - <X 2 / 


20. Extension of tbe Interval. The methods developed up to this point 
restrict the interval of expansion to (~7r,7r). In many problems it is de- 
sired to develop /(.r) in a Fourier series that will be valid over a wider inter- 
val. By letting the length of the interval increase indefinitely one may 
expect to get an expansion valid for all x. 

1 This theorem is due chiefly to Riemann. It is much deeper than the analogous state- 
ment for power series, and the proof would be quite out of place in the present book. 
See E. C. Titchmarsh, “The Theory of Functions,” pp. 427-432, Oxford University 
Press, London, 1950. 



INFINITE SERIES 


IS® INFINITE SERIES [CHAP. 2 

To obtain an expansion valid on the interval (~-2,Z), change the variable 
from x to fe/ir. If f(x) satisfies the Dirichlet conditions on (—2,2), then the 
function f(lz/v) can be developed in a Fourier series in z, 

0 do * A 

— h 2^ a n cos n 2 + 2^ b n sm nz (20-1) 

2 n—l n*»l 

for — t < 2 < v. Since 2 — t rx/l } the series (20-1) becomes 

Oq * nvx * mrx 

t fix) = 2^ a n cos — + 2 ^ Sin 

2 u «=1 2 n*l 2 


(20-2) 


By applying (18-4) to the series (20-1) , we see that 


«n 


1 r* fh\ 1 ri 

- I /I — ) cos nz dz = - I f(x) 

V \7 T/ l 


nvx 

cos dx 

l 


mrx 

dx -f 

f 2 

njrx \ 

1 

0 * cos 

/ 1 

* cos dx ] 

I rm — — 

i 2 

h 

2 / 

nr 

. mrx 

dx + 


nvx \ 

1 

0 • sin — - 

/ 1 

. sln 

~dx\ 

sm j 

5 2 

/o 

2 / 

nv 

1 , 2 / 

ITT 

1 

3ttj 

1 . 

&vX 


. mrx 2 
sm 

2 o 


1 o, 


1 r* (lz\ 1 ri nvx 

and b n * - / /I — } sin nz dz ~ - I f(x) sin dx. 

\7r/ l l 

As an illustration wc develop f(x) in Fourier series in the interval (—2,2) if f(x) *» 0 
for —2 < x < 0 and f(x) ** 1 for 0 < x < 2. Here 

-Kfj 
"•-Kfj 

Therefore, 

f 

If n is any integer, then 

ntr(x + 21) (nvx 

cos ~ cos — h 2mr 


• COS nir). 


; sm - 
5 2 


)■ 


) (nvi\ 

* 008 \ T/ 


and similarly for sines. Hence, each term of the series (20-2) has period 
22, and therefore the sum also has period 22. For this reason the sum can- 
not represent an arbitrary function on ( — it represents periodic func- 
tions only. 

Subject to the Dirichlet conditions, however, the function may be chosen 
arbitrarily on the interval (—2,2), and it is natural to inquire if a representa- 
tion for arbitrary functions on ( — qo,qo) might be obtained by letting 2 — » 

We shall see that such a representation is possible. The process leads to 



FOURIER SERIES 


SEC. 201 

the so-called Fourier integral theorem, which has many practical applica- 
tions. 1 

Assume that f(x) satisfies the Dirichlet conditions in every interval ( —1,1) 
(no matter how large) and that the integral 

M-f \f(x) | dx 

converges. As we have just seen, f(x) is given by (20-2), where 2 

1 f l nwt 1 ri nwt 

a n = 7 I fit) cos — dt, bn * ; / f(t) sin — dt. 

I 1 l l J—i l 


Substituting these values of the coefficients into (20-2) gives 


1 ri 1 * 

/(*) I dt + i^ I 


cos - 


mr(t — x) 

l 


dt 


when we recall that 


nwt utx nwt nwx nw(t — x) 

cos — cos j- sin — sin =* cos 

till l 


Since 



|/(j) | dx is assumed to be convergent, 


1 

21 



M 


<- C 1/(0 1 dt < 

2 l J -r ' 21 


(20-3) 


which obviously tends to zero as l is allowed to increase indefinitely. Also, 
if the interval ( — 11) is made large enough, the quantity w/l, which appears 
in the integrands of the sum, can be made as small as desired. Therefore, 
the sum in (20-3) can be written as 


1 ft 

- [A a I f(t) cos Aa(t — x) dt 
w J— 1 


AaJ f(t ) cos 2 A a(t — x) dt 

+ 

+ Aa f_ f(t) cos ?? Aa(t — x) dt 
+ ], (20-4) 

where Aa — w/l. 

1 Some of these applications are presented in Chap. 6. 

* We use t as variable of integration to avoid confusion with the x in (20-2). If /(a?) 
is discontinuous at x ** xo, the left side of (20-2) means ^[/(zo-f) 4* /(^o — )!• 



190 


INFINITE SEKIES 


[CHAP. 2 

This sum suggests the definition of the definite integral of the function 
F(a) = J *f(t) cos a(t — x) dt 

in which the values of the function F(a) are calculated at the points 

7T 2 7T 3 7T 

7 ~7' T' 

Now, for large values of l 

J f(t) cos a(t — x) dt 


differs little from 


J f(t) cos a(t — x) 

J — on 


dt 


and it appears plausible that as l increases indefinitely, the sum (20-4) will 
approach the limit 

1 r» rco 

-I da f(t) cos a[t — x) dt. 

71* • 0 * — oo 

If such is the case, then (20-3) can be written as 

f(x) = - r° da r f(t) cos a(f — x) eft. (20-5) 

7T 1 0 * — oo 


The foregoing discussion is heuristic and cannot be regarded as a rigorous 
proof. However, the validity of formula (20-5) can be established rigor- 
ously 1 if the function /(x) satisfies the conditions enunciated above. The 
integral (20-5) bears the name of the Fourier integral. 

Formula (20-5) assumes a simpler form if /(x) is an even or an odd func- 
tion. Expanding the integrand of (20-5) gives 


1 
7 r 



cos at cos ax dt + / /(/) 

J — oo 


sin at sin ax dt 




for the right-hand member. If /(/) is odd, then Jit) cos at is an odd func- 
tion times an even function, hence is odd. Similarly, f(t) sin at is even 
when/(0 is odd. Upon applying (19-4) to the first integral in the foregoing 
expression and (19-3) to the second integral, we see that 


/(*) 


2 

T 



sin at sin ax dt 


(20-6) 


1 See H. S. Carslaw, “Fourier’s Series and Integrals/’ pp. 283-294, The Macmillan 
Company, New York, 1921, or E. C. Titchmarsh, “The Theory of Functions/’ p. 433, 
Oxford University Press, London, 1950. 



me, 20] FOURIER SERIES 

when/(x) is odd. A similar argument shows that if f(x) is even, then 


191 


f(x) — d«j£V(0 cos at cos ax dt. (20-7) 


If f(x) is defined only in the interval (0 ,qo), then both (20-6) and (20-7) 
may be used, since f(x) may be thought to be defined in ( — <*>,0) so as to 
make it either odd or even. This corresponds to the fact that a function 
given on (0,ir) may be expanded in either a sine series or a cosine series. 

Since the Fourier scries converges to l /i\f{x+) + /(x — )] at points of dis- 
continuity, the Fourier integral does also. In particular for an odd func- 
tion 1 the integral converges to zero at x — 0, and this fact is verified by 
setting x = 0 in (20-6). 

Example. By (20-7) obtain the formula 

ir/2, if 0 < x < 1, 

ir/4, if X = 1, 

0, if x > 1. 

We choose /(x) * 1 for 0 < x < l and /(x) » 0 for x > 1. Then 

/** . . f x , sin a 

I J \ 0 cos at dt = / COR at dt * j a ^ 0. 

Jo Jo a 

Substitution into (20-7) gives 

f* 1 sin a r 

I COS ax da ~ - f(x) 

Jo a 2 

after multiplying by rr/2. Upon recalling the definition of /(x), we see that the desired 
result is obtained for 0 < x < 1 and for x > 1. The fact that the integral is *r/4 when 
x » 1 follows from 

1 r /0 + /0+) 

2 “ 2 


r sm 


sin « cos ax 


da 


PROBLEMS 

1. If /(x) is an odd function on (—/,/), show that the Fourier series takes the form 

V- , nVX , 2 [ l r, v N*\X , 

f(x) ~ 2^ sm — » b n » - / /Or) sin -7- dx. 

n—l l l Jo i 

Similarly, if /(x) is even, then 

. \ i X"' 7lJTX 2 TLirX 

f(x) « ““ + 2^ a n cos — > a n « 7 / /(x) cos — dx. 

* n*«l t l Jo l 

1 It should be noted that every odd function, if defined at x * 0, satisfies /(0) *» 0 

[although for an even function /(()) is arbitrary]. Hence a function defined for x » 0 
must sometimes be redefined at x » 0 before it can be made into an odd function. 



m 


INFINITE SERIES 


[CHAP. 2 


2* Expand the function defined by fix) * 1 on (0,2) &nd/(x) ** — 1 on ( —2,0). 

3. Expand /(x) * |x| in the interval (—1,1). 

4. Expand /(x) » cos *x in the interval ( — 1,1). 

5. Find the expansion in the series of cosines, if 

fix) « 1, when 0 < x < t, 

f(x) 0, when *• < x < 2r. 

Hint: Regard /(x) as being an even function. 

0. Expand 

f(x) ■» — x. if 0 < x < 

f(z) *■ x — if }i < x < 1 . 


7. Show that the series 


l • 1 2/iTTX 

- " «« ~r ' 

7r narl n l 


represents x Al — x when 0 < x < L 

8. Show, with the aid of (20-6) and (20-7), that 

r at sin ax , ir „ 

«£* « - e^ 1 , if ft > 0, 

. a 2 + /S 2 2 

r cos ax , 7T 

Hint: Take /(x) = 

8. An integral equation is an equation in winch an unknown function appears under 
an integral sign. If F(t) is known and fix ) is to be found, the integral equation of F ourier is 


I fix) cos xt dx ** F(t). 

J o 


(a) Using (20-7) show that a solution is given by 


fix) — - j Fit ) cob xt dt. 

T Jo 


ib) State a similar integral equation which can be solved by use of (20-6), and 
solve it. 

21. Complex Form of Fourier Series. The Fourier series 


u o 

f(x) =* f- 2Lf ( a n c; os nx + b n sin nx) 

2 n* 1 


with a n — - f f(t) cos nt dt , K ~ ~ f /( 0 si 

7T * 7T ' * 

can be written, with the aid of the Euler formula 1 


sin nZ dt 


e lu « cos u + i sin n 


( See Sec. 17. 



SBC. 21] FOURIER SERIES 

in an equivalent form, namely, 

/(*) - E Cn& in *t 

<0 

where the coefficients c n are defined by the equation 

e„ = ^ r f(t)e~ int dt. 

2r •'-* 


193 

(21-3) 

(21-4) 


The index of summation n in (21-3) runs through the set of all positive and 
negative integral values including zero. 

The equivalence of (21-3) and (21-1) can be established in the following 
manner: Substituting from (21-2) in (21-4) gives, for n > 0, 


c n ~ — r os nt — i sin rd) dt 
2 ir 

1 pr i rr 

-- — / /(*) cos nt dt / /($) sin eft 

2tt ■'-* 27r ■'-* 


.&n 

¥ 1 "2** 


A similar calculation gives 


. b« 

C_ n = 

2 2 


Oo 

2” 


while c 0 

Now (21-3) can be written in the form 

oc 00 

/(*) = c 0 + E c n c ,MI + E c_ n c- <n *. 

n=aal — 1 


n* 1 


Making use of the expressions for the c„ just found gives 


fix) 


a 0 . ^ a n ~ ib n wx 


e- + £ 

* n=»l 


2 + E 

•* n ~1 

o„ e’ ni + e~~' nx 

- + E «« x 

“ n as 1 “ 


On + ?'&n 


~ ' E & n 

n«*l 


»*** _ ~ — tnx 


By (21-2), 

e* w _|~ a* 2 cos u and e tu — e~ tu = 2f sin t* 
and hence the latter series is identical with (21-1). 



194 INFINITE SERIES [CHAP. 2 

To illustrate the use of (21-4), consider the function f(x) » e ax on (— ir,ir). Here, 


2rrc, 


-r, 


f" t e ~,n( dl 


■ f e‘“" 
J — r 


">‘dt 


e (a— w)t jir 


_ e ~ 


a in \ -x 

Since (21-2) gives e ±tn * * cos (±mt) ** (*-l) n , we obtain 

(-l) n SinllTra (~l) n 


c n « 

and hence by (21-3), 


a — m 
smh ira 


~ Tj" 2 ^ + m ) 
-f tr 


ir n «.--«o + W 

The methods of the last section yield 


£ „8 + i ' ! ) £,m:c - 


(21-5) 


/(sr) = E c n e m * xl ‘, with r„ = ~ dt, (21-0) 


for the expansion on an arbitrary interval ( — Z, /) Upon letting Z co, we 
obtain the Fourier integral theorem in the form 


A*) * lim ~ f da f f(()e 

A * 2tt J ~ A J — « 


,ta(z— 0 


(ft 


when/(x) satisfies the conditions postulated for (20-5). 
If 


0(«) = -4= f ^ lw 7(^) 

V27r y — 00 


dr, 


then (21-7) gives, after renaming some of the variables, 

/•A 

a * \/2i r 

The transform T defined by 


1 [A 

/(x) = lirn — / F ux g{u) du . 

a / 0*rr J — ' 


(21-7) 


( 21 - 8 ) 


(21-9) 


t(/) - ~ 4 r r e ~' ux l^ dx 

V2t 


is called the Fourier transform; it is one of the most powerful tools in the 
whole repertoire of modern analysis. Although T is related to the Laplace 
transform L introduced in Appendix B, T is much easier to invert; that is, 
one can readily find A#) by (21-9) when T (J) is knowm 


PROBLEMS 

1. Derive (21-6) from (21-3) and (21-4) 

2 . (a) Show that 

Oq » 2C0, a n * Cn + &« “ *(C « — C-~n) 

[and hence the real form (21-1) can be deduced from the complex form (21-3)]. 
( b ) By applying your result (a) to (21-5) obtain the real Fourier series (21-1) for e**. 



SEC. 22] ADDITIONAL TOPICS IN FOURIER SERIES 

3. By setting x « 0 in (21-5) obtain the expansion 

v ^ <-l) n 

— - as y — - .. 

a sinh Ta n «— «> 4* n 2 


ADDITIONAL TOPICS IN FOURIER SERIES 

22. Orthogonal Functions. A sequence of functions B n (x) is said to be 
orthogonal on the interval (a, b) if 


f 6 m (x)e n (x) 

Jn 


for m 5 * n, 
for m = n. 


r B m (x)B n (x) dx = f sin mx sin nx dx = 

It r/: 


For example, the sequence 

0i(x) « sin x, 0 2 (x) = sin 2x, . . B n (x) = sin nx, 

is orthogonal on (0,tt) because 

r r* . (0 for m t* n y 

B m (x)Q n (x) dx = / sin mx sin nx dx = 

•'o l tt/2 for m = n. 

The sequence 

1, sin x, cos x, sin 2x, cos 2x, ... (22-2) 

is orthogonal on (0,2x), though not on (0,7r). 

In the foregoing sections the functions (22-2) were used to form Fourier 
series. Actually, one may form series analogous to Fourier series hv means 
of any orthogonal set. These generalized Fourier series are an indispensa- 
ble aid in electromagnetic theory, acoustics, heat flow, and many other 
branches of mathematical physics. 1 

The formula for Fourier coefficients is especially simple if the integral 
(22-1) has the value 1 for m -= n. The functions B n (x) are then said to be 
normalized, and {0 n (x)| is called an orthonormal set. If 


f [0 K {x)f , 


in (22-1), it is easily seen that the functions 

4>nM = (4 n)~^B n (x) 


are orthonormal; in other words, 


rb 

I )<f> n (x) i 

Ja 


= 0 for m ^ n, 

= 1 for m = ft. 


For example, since 


r2v r‘2* /*2ir o 

/ 1 dx - 2tt, / sin 2 nx dx ~ 7r, / cos“ nx dx =» tt 

Jo Jo Jo 


1 See Chap. 6. 



INFINITE SERIES 


m 


[chap* 2 


for n > 1, the orthonormal set corresponding to the orthogonal set (22-2) is 
(2**)"“^, tt~ h sin x, cos x, , . sin nr, cos nx, 


The product of two different functions in this set gives zero but the square 
of each function gives 1, when integrated from zero to 2ir, 

Let {<f> n (z) } be an orthonormal set of functions on (a, 6), and suppose that 
another function /(x) is to be expanded in the form 

J{x) = Ci4>i(x) + c 2 <t> 2 (x) -f f* Cn4>«(*) 4 . (22-4) 

To determine the coefficients c n we multiply by <t> n (x), getting 

f(r)<t>n(x) = c y <t> i(x)0n(x) H f c n [<A„(x)] 2 H . 

Here, the terms not written involve products 4> n (x)4> m (x) with rn ^ n. If 
we integrate from a + o b , these terms disappear, and hence 

f f(x)<f> n (x) dx ~ f c n [<j> n (x)] 2 dx = c n (22-5) 

-/a Ja 


According to Theorem III, Bee 7, the term-by-term integration is justified when the 
series is uniformly convergent and the functions are continuous The foregoing pro- 
cedure shows that if f(x) has an expansion of the desued type, then the coefficients r n 
must be given by (22-5). In the following section (22-5) is obtained in a different manner, 
which does not assume uniform convergence 


The formula (22-5) is called the Euler -Fourier formula, the coefficients 
c n are called the Fourier coefficients of /(.r) with respect to {</>„(x)J, and the 
resulting series (22-4) is called the Fourier senes of f(x) with respect to 
{ <f> n ( K x ') } . The reader can verify that the foregoing results applied to the 
sequence (22-2) >ield the ordinary Fourier series, as described in the fore- 
going sections. 

Orthogonal sets of functions are obtained m practice by solving differen- 
tial equations, and this possibility will be discussed next On a given inter- 
val a < x < b consider the equation 


d 

dx 



+ g(*)y = M*)y, 


A ~ const, 


(22-6) 


or, in abbreviated form, 

(py'Y 4- qy - \ry, 


d 

d~x 


It will be convenient to require the additional condition 

f ry 2 dx j* 0 


which, in particular, rules out the trivial solution y ^ 0. 



SEC, 22] ADDITIONAL TOPICS IN FOURIER SERIES 197 

Let y m be a solution when X has the value \ m , and let y n be a solution 

when X has a different value, X n . Thus, 

(py'mY + QVm “ X m rj/„„ (22-7) 

(.py'nY + qy n = x„n/„. (22-8) 

If (22-7) is multiplied by y n and (22-8) by y mi we get 

Vnipi/mY - VmipynY * X w r// m it/ n - \ n ry m y n (22-9) 


after subtracting the resulting expressions. Since 

fl/*.(W«) ^ ?/m(?Vn)]' = VnipUmY 4* IfnWm) ~ VmipyhY 
=• left side of (22-9), 

the foregoing result (22-9) may be written 
rf 

*T~ [p(2/n^/m 2/m?/n)] ^ (X*n X n )n/ w 2/ n . 

ax 


y’mivy'n) 


Integrating from a to b yields the fundamental formula 



when r is continuous. 

If the conditions at a and b are such that the loft side of 
we can deduce 


( 22 - 10 ) 


(22-10) is zero. 


/: 


Wml/n djT = 0, 


m n, 


( 22 - 11 ) 


since X m ^ X w . The relation (22-1 1) may be written 

( ( V'r (/,„)( V r y n ) d.r = 0, m 9 * n, 

■fa 

and hence the sequence d u (x) defined by 

^ l/n * vV(x)2/ w (x) (22-12) 

satisfies the orthogonality criterion (22-1 J. An orthonormal set {<£ n j may 
be obtained from {0 n j as described previously. 


When r(x) is negative, the foregoing prows does not yield a real sequence { 0 n (x ) } , and 
it is better to work directly with (22-11). Functions y n satisfying (22-11) are said to be 
orthogonal unth respect to the weighting function r(x) , the definition (22-1) corresponds to 
the case r be 1. Fourier series based on the more general concept of orthogonality (22-1 1) 
are quite analogous to those based on (22-1) (cf. Frob 2) 

Example l Show’ that the sequence (22-2) is orthogonal on the interval ( — t,*-). 
Since sin nx and cos nx satisfy y" *=* - n l y, we may use the formula (22-10) with 
p»r» 1, The result is 


(VnV'm ~ VmVn) 


( — to 2 -f* n 2 ) f y m y n dx, 

• — T 


(22-13) 



198 


INFINITE SERIES 


[chap. 2 


where y n » sin nx or cos nx and y m « sin mx or cos mx. Since y % Vm — has period 
2*, the value at % is the same as the value at —rr, and hence the left side of (22-13) is 
zero. This yields the desired orthogonality except in the case m « n. If ?n « n, how- 
ever, the relevant integral may be evaluated by inspection: 


/: 


cos nx sin nx dx 



sin 2 nx dx 


0. 


Example 2. Show that the Legendre functions P n (x) are orthogonal on the inter- 
val (-1,1). 

Legendre’s equation (13-8) may be written 

[(1 - *V)' - xp, 

where X is constant; X * — n(n -f- 1) when y » P n ( x). The special case p «* (1 — x 2 ), 
q o, r « 1 in (22-0) and (22-10) yields 

(1 - x 2 )(P n P; - P m K) f - f — m(m + 1 ) + «(w + D1 I' PmU)Pn(Jr) dx 
l-i 

Since (1 — x 2 ) vanishes at dbl, the left side is zero and the oithogonahty follows. It can 
be shown 1 also that ^ 

f IP„(r)] 2 dx = — T”7’ 

2n -f 1 


and hence the corresponding orthonormal set is 

<Pn(x) * (« • f 1 2)‘^Pn(x) 

Example 3. l*et the sequence p\, p 2 , • - • be the distinct positive roots of the equation 
Jp(x) « 0, so that J^Pn) =* 0. If ju > 0 the functions 


4>n(x) 


v-r/ ,JpnT) 


are orthonormal on the interval (0,1). 

By (14-1) it is found that y — J^ipx) satisfies 


J'M) 


( 22 - 1 1 ) 




m 2 v 


Xxy, 


and hence (22-10) holds with p * r » x. If we clioose » J>(px) and t/ m ** ^(pmx), 
the left side of (22-10) is 

\ d d l I 1 

x L ./ M (PX) J p(pmX) J p(pmX) ^~Jp(pX) J | , 

which reduces to JpijfipmJlSpm), since J»(p m ) * 0. It follows that 


(~~Pm ~f~ p 2 ) f xJ,(p m x)JM) dx «* PmJ p(p)J n(pm) • 
J 0 


(22-15) 


1 See E. J. Whittaker and G. N. Watson, “Modern Analysis/’ p. 305, Cambridge 

University Press, London, 1952; J. M. Macltobert, “Spherical Harmonics/’ p. 92, Dover 
Publications, New York, 1948; W. E. Byerly, “Fourier Series and Spherical Harmonics/’ 
p. 170, Ginn & Company, Boston, 1893. 



SRC. 22 ] ADDITIONAL TOPICS IN FOURIER SERIES 

Since / M (p n ) ** 0, the choice p « p„ in (22-15) yields 


I Xj p{pm^)J n{.PnX) (lx — 0, 71% 5^ 71. 

JO 


Moreover, differentiating (22-15) with respect to p we get 


2p f xJ dx *j~ (p 2 Pm) f X~J n(p m x)J ^(px) dx PmJ fi{p)J JjiTn) , 

Jo Jo 

which reduces to l 

2 [ x[J,( Pm x)f dx - f/;(p m )] 2 ( 

J o 


when p » p m . Equations (22-16) and (22-17) show that the sequence (22-1 -1) is ortho- 
normal on (6,1), as desired. 

The fact that the equation / M (x) ■= 0 has infinitely many roots p n is established in 
treatises on Bessel functions; analysis of such questions for general differ entail equations 
constitutes the so-called Sturm-Lio wnlle theory. It can be shown that Fourier senes of 
Bessel or Legendre functions actually converge; that is, an analogue of Diriehlet’s the- 
orem holds in such cases. 1 These questions are treated, from a very general point of 
view, m a branch of analysis known as spectral theory. 


PROBLEMS 

1. By considering the equation y" * \y show that the sequence sin riwx ! l is orthogonal 
on the interval (0,0, and construct the corresponding oithonormal sot. 

2. Suppose an arbitrary function f{x) is expanded in a uniformly convergent series 
f(x) « wC n t/ 7 iW, whore y n are the functions m (22-11) Show that 

Cn ** ^ r VMyn(jr) dx^ r(x)\y n (x)] 2 dx^ . 

Hint Multiply the given senes by i{x) y ti {x), and integrate term by term. 

3 . If m -J- n is positive, show that. 

(m 2 - rr) f x~\r tn {x)Ju(x) c lx « D ~ Sm(l)Jn(l)l 

Jo 

Hint * Bessel's equation (1 1-1) may be written 

X 

{xy ) + xy = - y, 
x 

where X « n 2 when y — JJx) To avoid difficulty at x * 0 one may consider j and 
let e —» 0. The convergence follows from (1 1-0), since (11-9) gives 
Jm{x)J n (x) ^ (const) x m+n , an x — » 0 

4 . It can be shown that as |/| — ♦ 


[J 

/ 7 r 7ir\ 

J&) ~ 


1 2 . ( 

it m r\ 

J— COS 1 

V ra* 

V ~ 4 ~ TJ ' 

j — * 

Cr m v 

1 

»■* i 

1 


1 E, A. Coddington and N. Levinson, “Theory of Ordinary Differentia] Equations,” 

chap. 7, McGraw-Hill Book Company, Inc.. New Xork, 1955 



200 INFINITE SERIES [CHAP, 2 

By letting / « in Prob. 8, deduce 

r 2 r 

x~ ] l J m (x)J n (x) dx * - sin (rn — ») ~ > to 4- n > 0. 

7T 2 


6 . If <f>n(x) are orthonormal on the interval (0,1), show that 

^r«W * 0“ ! Vn(x/0) 


are orthonormal on the interval (0,n) 


23. The Mean Convergence of Fourier Series. If we try to approximate 
a function /(x) by another function 7 ? n (.r), the quantity 

!/(*) - />n(x)| or [/(x) - p„(x)] 2 (23-1) 


gives a measure of the error in the approximation. The sequence p n {x) 
converges to f(x) whenever the expressions (23-1) approach zero as n — > 00. 

These measures of the error are appropriate for discussing convergence 
at any fixed point x. But it is often useful to have a measure of error which 
applies simultaneously to a whole interval of x values, a < x < b. Such 
a measure is easily found if we integrate (23-1) from a to b: 

f !/0) - PnMi dx or f f f{x) - p„(x)] 2 rfx. (23-2) 

Ja Ja 


These expressions are called the ?nean 1 error and mean-square error , respec- 
tively, If either expression (23-2) approaches zero as n co, we say that 
the sequence p„(j) converges in mean to f(x) and we speak of mtan con - 
vergerwe . 

Even though (23-2) involves an integration which is not present in (23-1), 
for Fourier series it is much easier to dbcuss the mean-square error and the 
corresponding mean convergence than the ordinary convergence. Such a 
discussion is presented now. 

Let 4>„{x) be a set of functions normal and orthogonal on a < x < b, so 
that, as in the last section, 


f h 

I 10m (*r) dx 

Ja 


0 for m n f 

1 for m = n. 


(23-3) 


We seek to approximate f(x) by a linear combination of 0„(x), 

Pn(x) * «i0i(t) 4- a 2 </> 2 (x) 4 b a n 4>„(x), 

in such a tvay that the mean-square error (23-2) is minimum: 2 

1 Note that if the expressions (23-2) are multiplied by 1/(6 — a), we get precisely the 
mean values of the eoi responding expulsions (23-1). 

4 We use / and <f> n as abbreviations for f(x) and <J> n (x), respectively. It is assumed that 
/ and <t> n are integrable on a < x < 6. If the integrals are improper, the convergence of 

fb rb 

I / 2 dx and / ef>% dx is required. 

Ja Ja 



sec. 23] 


ADDITIONAL TOPICS IN FOURIER SERIES 


201 


rb 

& m J If ~ (°i*i H h «n *«)] 2 dx « min, (23*4) 


Upon expanding the bracket we see that (23-4) yields 

f f 2 dx — 2 f 4 h a n <£ n ) dr + f (a^ -j |~ a n 4> n ) 2 dx . 

•/a •'a Jq, 

If the Fourier coefficients of / relative to <#> fc are denoted by 

Ck = ( f<t>k dx, 

J a 

then the second integral (23-5) is 


(23-5) 


rb 

I /(^1</>1 + * * ' + ^n4>n) dx = dyCi + a 2 C 2 + * * * + a n C n . 

J a 

The third integral (23-5) may be written 

rb 

j (d\<t > i + * * • + &n<£n)(& 1^1 + • * * + O'n^n) dx 
Ja 

= (af 0i + + * — h a 2 tf> 2 + • * • ) dx 

= af +■ * • * -f- a 2 , 

where the second group of terms “H — > ” involves cross products <f>^ ; with 
i 7* j. By (23-3) these terms integrate to zero, and the expression reduces 
to the value indicated. 

Hence, (23-5) yields 

E s f f dx - 2 2 a k c k + £ (23-6) 

a A*- 1 k sw 1 


for the mean-square error in the approximation. Inasmuch as 
-2 a k c k + a* s -cj[ + (a* - c*) 2 , 
the error E in (23-6) is also equal to 

£ = /V 2 dx - £ cl + £ (a* - c t ) 2 . (23-7) 

*' rt A- l 

and we have established a theorem of central importance: 

Theorem I. If <p n Is a set of normal and orthogonal functions , the mean- 
square error (234) may he written in the form (23-7), where c k are the Fourier 
coefficients of f relative to 4> k - 

By going back and forth between the two expressions (234) and (23-7), 
one obtains a number of interesting and significant theorems with the 
greatest ease. In the first place, the terms ( a k — c k ) 2 in (23-7) are positive 



INFINITE SEBIE8 


202 


[chap. 2 


unless dk « Ck , in which case they are zero. Hence the choice of a k that 
makes E minimum is obviously a * = c kf and we have the following: 
Corollary 1 . T/ic partial sums of the Fourier series 


Ci4>i + • — h c n (j> nj 
rb 


Ck 


■ Jj4>kdx 


ro 

give a smaller mean-square error / (/ — p n ) 2 dx than is given by any other 

J a 

linear combination 

Pn ~ -{“••• -f- a n <t> n > 


Upon setting a k = c k in (23-7), we see that the minimum value of the 
error is 


min 



n 


Ed 


(23-8) 


Now, the expression (23-4) shows that E > 0, because the integrand in 
(23-4), being a square, is not negative Since E > 0 for all choices of a k , 
it is clear that the minimum of E (which arises when a k = c k ) is also > 0 
The expression (23-8) yields, then, 


( f 2 dx - > 0 or T.4< ('f 2 dx. 

Ja k~* 1 k—A Ja 

Upon letting n — * awe obtain 1 

fb 

Corollary 2. If c k ~ I f<t> k dx are the Fourier coefficient s of f relative to the 

Ja 

orthonormal set <t> H then the series LYjr converges and satisfies the so-called Bessel 
inequality 

CC b 

E 4 < / f/Wf dx. (23-9) 

*~1 J<1 

Because the general term of a convergent series must approach zero 
(Sec. 1, Theorem I) we deduce the following from Corollary 2: 

fb 

Corollary 3. The Fourier coefficients c n — / /<£ n dx lend to zero as n -- ► oo. 

J a 

For applications it is important to know whether or not the mean square 
error approaches zero as n oo. Evidently the error approaches zero, for 
some choice of the a k s, only if the minimum error (23-8) does so. Letting 
n — > oo in (23-8) we get the so-called Parscval equality 

f f dx - E c* = 0 
Ja 

as the condition for zero error: 


1 Since c* > 0, the sequence 2 r l is nondecreasing. We have just seen that it is 
bounded, and hence the limit exists by the fundamental principle (Sec. 1). 



ADDITIONAL TOPICS IN FOURIER SERIES 


203 


sec. 23] 

Corollary 4. Iff is approximated by the partial sums of its Fourier series, 
the meaursquare error approaches zero as n — ► <x> if f and only if, BesseVs 
inequality (23-9) becomes ParsevaVs equality 

23 c l = f [/(• r ) ] 2 dx. (23-10) 

n=t Ja 

In other words, the Fourier series converges to / in the mean-square 
sense if, and only if, (23-10) holds. If this happens for every choice of /, 
the set 4> n (x) is said to be closed . A closed set, then, is a set that can be 
used for mean-square approximation of arbitrary functions. It can be 
shown that the trigonometrical functions cos nx and sin nx are closed on 
0 < x < 2t r, though the proof is too long for inclusion here 1 

A set <t> n (x) is said to be complete if there is no nontrivial function 2 
f(x) which is orthogonal to all of the <£ n s. That is, the set is complete if 

Ck * f f(x)<l>k(x) dx - 0 for k = 1, 2, 3, . . (23-11) 

implies that 

A/(t)] 2 dx « 0. (23-12) 

J a 

Now, whenever (23-10) holds, (23-11) yields (23-12) at once. Hence we 
have : 

Corollary 5. Every closed set <p n (x) is complete. 

The converse is also true Every complete set is dosed This converse, However, requires 
a more general integral than that of Ricmann The generalized integral is known as 
the Ijebesgue integral; it was first constructed to deal vwtli this very pioblem. A brief 
description of the Lebesgue integral is given m Appendix C 

The notions of closure and compUU tui ss have simple analogues in the elementary the- 
ory of vectors Thus, a set of vet tots Vi, Vi, Vg is said to be dosed if every vector V can 
be written in the form 

V - fiVi + r 2 V 2 + mV* 


for some choice of the constants c k The set of vectors Vj, V 2 , V 3 is said to be complete 
if there is no nontrivial vector oithogonal to all of them That is, the set is complete if 
the condition 

V-V* - 0 for * « 1, 2, 3 

implies V*V ** 0. 

In this setting, it is obvious that closure and completeness are equivalent, for both 
conditions simply state that the three vectors Vj, V 2 , Vg are not coplanar. These matters 
are taken up more fully in Chap. 4. 


1 See E. O. Titchmarsh, “The Theory of Functions," p. 414, Oxford University Press, 
London, 1950 

* In the theory of mean convergence fix) is regarded as trivial if f(x) * 0 for so many 


values of x that 



dx 


0 . 



204 


INFINITE SERIES 


[CHAP. 2 


PROBLEMS 

1. (a) Show that Parse va Is equality takes the form 

1 r 2V 1 oo 

- I (/(*) 1 2 <ir- 50 o + Z(<4 + 6 J) 

» •/» 2 „_i 

when ^»(x) are the trigonometric functions on (0,2r). ( b ) Specialize to sine and cosine 
series on (0,ir). 

2. It is desired to approximate 1 by 

p(x) ** ai sin x 4* as sin 2 j + sin 3x 

in such a way that J fl — p(x)] 2 dx is minimum. How should the coefficients a % be 
determined? 0 

3. Give a direct proof that as n — + oo, 

rt* r2* 

I f(x) sin nxdx -+ 0, I f(x) cos nx dx — > 0, 

J o Jo 

if f(x) is periodic of period 2 ic and has a continuous derivative f’(x ). Hint: Integrate 
by parts, 

4. Obtain the formula a* ** c* from (23-1) by using the fact that dE/dak * 0 at the 
minimum value of E . 

24* The Pointwise Convergence of Fourier Series. We shall now obtain 
an explicit formula for the difference between a function and the nth partial 
sum of its (trigonometric) Fourier series. The formula will enable us to 
establish the convergence for a class of functions which includes all the 
examples given in this book. 

If f(x) is a bounded integrable function of period 27r, the nth partial 
sum of its Fourier series is 

1 n 

s n (x) « ~ a 0 + X) ( a k cos kx + b k sin fee), (24-1) 

2 k^i 

where the coefficients are given by 

a k = - f f(t ) cos kt dt, = - [ f(t) sin kt dt. (24-2) 

IT 7r — T 

Substituting (24-2) into the series (24-1) we get 



SBC. 24] ADDITIONAL TOPICS IN FOURIER SERIES 

If we define the so-called Dirichlet kernel by 

3 n 

D n (u) ** - + £ cos ku, 

2 hum l 

the foregoing result takes the simpler form 

*»(*) = - f f(f)D n {i - x) dt. 

7T — W 

Setting t — x ~ u in (24-4) yields 

1 fit —X 

s n (x) = - / /(/ + u)D n {u) du. 

7T 


205 

(24-3) 

(24-4) 

(24-5) 


Now, 2> n (u) has period 2ir by inspection of (24-3), and /(r) also has period 
2 7T. Hence, the integral of /(u + x)D„(u) over any interval of length 2w is 
the same as the integral over any other interval of length 2n, and (24-5) 
may be replaced by 

1 fT 

s n (x) « - / f{x + u)I) n (u) du . (24-0) 

71- ■/ ~ir 

Since D n (~~u) = D n (a) by (24-3), we may replace u by —a in (24-6) to 
obtain the alternative form 

1 f* 

«»Cr) = - / /(* — u)D n (u) du. v 24-7) 

7T 7T 

The sum of (24-6) and (21-7) yields 

2* n (-r) * - f (/(-r + a) + /(-r - i/) 1A» du. 

7T ^ — Jr 

Since the integrand is an even function of u, the integral from 0 to w is half 
the integral from — t to w, and we have thus established that 

t f* 

« n (x) ® ~ / lf(x + u) + fix - u)]D n (u) du. (24-8) 

7 r J() 

To introduce /(x) into our considerations, we observe that 

~ « - fD n (u) du, (24-9) 

2 7T "'C 

since the terms involving cos ku in (24-3) integrate to zero. If (24-9) is 
multiplied by 2 f(x) (which is constant with respect to the integration var- 
iable u), we get 

f{x) = - f 2 f(x)D n (u) du . 

7T •'0 


(24-10) 



20(5 INFINITE SERIES [CHAP. 2 

Subtracting (24-10) from (24-8) gives the fundamental formula 

s„(x) - f(j) - 1 jT; [/(x + M) - 2/(x) + /(* - u)}D„(u) du, (24-11) 

which will now be used to study the convergence of s n (x) to f(x). 

We shall say that f(x) is 'piecewise smooth if the graph of /(.r) consists of 
a finite number of curves on each of which f'{x) exists. We suppose also 
that the derivative exists at the end points of these curves, in the sense 


f(x + u) - f(x+) 

lim — 

u — » 0 + u 


or 



/(X - u) - fi x-) 
— u 


(24-12) 


where “u — » 0+” means u —> 0 through positive values. Such a function 
may have finitely many discontinuities. However, since the Fourier co- 
efficients of f(x) are not altered if f{x) is redefined at a finite number of 
points, we can assume that 


/to = 


/(.r-f )+/(*-) 
2 


(21-13) 


at every point x, whether /(x) is continuous at x or not. 

These preliminaries lead to the following theorem: 

Theorem. If f(x) ts periodic of period 2ir, is piecewise smooth , and is 
defined at points of discontinuity by (24-13), then the Fourier series for /(. r) 
converges to f(x) at every valm of x. 

To establish this theorem we recall that the series (24-3) was summed in 
Sec, 17, Example 1. The result (17-5) yields 


D n (u) 


sin (n + 
2 sin 


(21-11) 


If we substitute this into (24-11) and replace 2/(.r) by f(x+) + /(•** — ) in 
accordance with (24-13), we get 


«»(*) - f(x) 


i r /(* + u ) - /(*+ )+/(*-«)- f(x~) . 


= -/ 

IT J 0 


sin 


2 sin 

Now, the expression 

fix + U ) - f(xj-) _ f(x + u) - /(x+) (m/2) 

2 sin w 

has a limit as u — > 0+, since 


u du. 


sin (r/2) 


(24-15) 


uj 2 
lim 


1 


sin (r/2) 


as w 0 



SEC. 25] ADDITIONAL TOPICS IN FOURIER SERIES 207 

and since the limits (24-12) exist by hypothesis. If we define the value of 
(24-15) at u * 0 to be this limit, then the expression is continuous for 
u > 0 as long as the points 

[x,/(x+)] and [x + u , f(x + u)] 

are on the same smooth curve belonging to the graph of /(x). 

On the other hand, for u ^ 0 the function (24-15) is just as well behaved 
as the numerator f(x + u) — f(x+), since sin y 2 u does not vanish. This 
shows that the graph of (24-15) consists of a finite number of continuous 
curves, which have finite limits as one approaches their end points and 
hence are bounded. 

It follows from Corollary 3 of the preceding section 1 that 



f(x + u) - f(x+) 
2 sin y 2 u 



udu ~ 0 . 


In just the same way it is found that 


lim 

n — + « o 


I 


- u) 


2 sin y 2 u 


■/(*-> 

— sm 


in u du = 0 


and hence tin 1 integral representing s n (x) — f(x) tends to zero as n 
This shows that 


lim s n (x) = /( x) 


n --*»«• 


00 . 


and completes the proof of the theorem. 

26. The Integration and Differentiation of Fourier Series. If f(x) is 
piecewise continuous 2 on [ — 7 r ? 7 r], then the function 

m -[mat ( 25 - 1 ) 

is continuous and piecewise smooth (Sec. 24). Moreover, F(x) remains 
continuous when defined to have period 2*, provided F(~ tt) = F(tt). 
Since F(~-w) = 0, the latter condition reduces to 

F(t) = dt = jtOo = 0 (25-2) 

where l / 2 a 0 is the first Fourier coefficient of /(x). Applying the theorem of 
1 The presence of the Vi in sin (n + M)u causes no trouble, since 
sin (n 4* 1 <i)u — sin nu cos x /%u 4* cos nu sin 
and Corollary 3 applies to each term. 

*This means that the interval [-tt,*] ran be divided by points xi, x% . . x n into a 
finite number of intervals on each of which f(x) is continuous. Also f(x) must have a 
limit as x — ► £*4- and as £ — * 



208 INFINITE SERIES [CHAP. 2 

the preceding section, we can now deduce that the Fourier series for the 
periodic function F(x) converges to F(x) at every value of x. 

This result can be obtained when f(x) and \f(x) ) are only assumed integrable, without 
being piecewise continuous. Indeed, we can always write f(x) *» P(x) — N(x) where 
P(x) m positive and — N(r) is negative. The equation 

F(x) « r P(t) dt - r N(t) dt 
J — r J -~r 

expresses F(x) as the difference of two increasing continuous functions. Since such 
functions satisfy the Dirichlet conditions, the desired result can be deduced from Diri- 
chlet’s theorem as quoted in Sec. 18. 

We shall show next that the Fourier series for F(x) is obtained by inte- 
grating the series for fix). If n > 1, the Fourier cosine coefficient A n of 
F(x ) satisfies 

r* sin i 

TrA n — / F(x) cos nx dx = F(x) - — 

J ~* n 

when we integrate by parts. Since F( — t) = F(tt) * 0, the integrated 
part drops out, and since F'(x) — fix), the expression becomes 

1 rr b n 

jrA n = / sin nx fix) dx = —w — 

n J —* n 

In the same way B n — a n /n , and also 

1 f* 

A 0 ~ I xf(x) dx. (25-3) 

v * 


rr sin nx 

- / F’(x) dx 

J —t n 


These considerations establish the following remarkable theorem: 

Theorem I. Let fix) be a function of period 2w which has a Fourier series 

2(a n cos nx + b n sin nx). (25-4) 

Then, with A 0 given by (25-3), 


f m 

J — r 


dt 


1 


; -4o + 2 


/^n 

\n 


sm nx cos nx 

n 


)■ 


(25-5) 


and this equation holds for all x , even if the Fourier series (25-4) does not 
converge. Moreover , the series (25-5) is actually the Fourier series of the 
function on the left. 

In case a 0 ^ 0, so that the Fourier series for fix) is 
^a 0 + 2(a* cos nx + b n sin nx), 
we apply Theorem I to fix) — 3^o 0 . Inasmuch as 

ff{x) dx - [ $ fix) dx - P fix) dx * Ftf) - Fia) 

J Ct J * — — y 



209 


BBC. 26] ADDITIONAL TOPICS IN FOURIER SEHIBS 

lor all a and /S, the reader may deduce, by Theorem I, that 
fP rP rP 

I f(x) dx = I (J^Oo) dx + 2 / (a n cos nx -f b n sin nx ) dx. (25-6) 

Ja Ja J<x 

This result may be summarized as follows: 

Theorem II. Any Fourier series {whether convergent or not) can be in- 
tegrated term by term between any limits . The integrated series converges to 
the integral of the 'periodic function corresponding to the original series. 

For example, according to (18-5) the Fourier series for ]/%% is 


*r x = sin x 
2 


sin 2x sin 3x 

j 




(26-7) 


If we integrate from a to x by Theorem II, we get 


1 , * 2n A , cos nx — cos na 

~(x i - a 2 - Z(-l) n j 

4 n- 1 n 2 


•** - C + £( -1)" 


Treating a as constant, we see that 
1 

4 n"^i n* 

where Cis constant. Since C is the first Fourier coefficient of 

r 1 

2ir , 

Alternatively, because do * Om (25-7), we can use (25-3) to obtain 


and hence, by (25-5), 

F{x) 


l r* l 
2k L, 4 
(25-7), we 


x 2 dx 


K 

u 


l 


2 ) 


x 2 dx 




(25^) 


(25-9) 


4 ' 6 n-i n* 

The consistency of this result with (25-8) and (25-9) is easily verified. 

Although Fourier series can always be integrated, as we have just seen, 
the differentiation of Fourier series requires caution. For example, the 
series (25-7) converges for all x, and yet the series 

cos x — cos 2x + cos 3x — cos 4x H 


obtained by differentiating (25-7) diverges for all x. The trouble is that 
the function x /ix (when made periodic) has no derivative at the points =tw, 
=fc3x, dt 5ir, 

This example is quite typical of the general situation, which can be de- 
scribed as follows: There is not much hope of being able to differentiate a 
Fourier series , unless the periodic function generating the series has a deriva- 
tive at every value of x. On the other hand, when this condition is fulfilled, 
we usually can differentiate, as is shown by the following theorem; 



210 


INFINITE SERIES 


[CHAP. 2 

Theorem III. Let f(x) have period 2r, and suppose f'(x) exists for every 
value of x, without exception. If f(x) is continuous, 1 the Fourier series for 
f(x) can be obtained by differentiating the Fourier series for f(x). If f{x) 
is continuous and has only a finite number of maxima and minima on [~ 7 r, 7 r], 
the differentiated scries actually converges to f(x) for every x. 

Repeated application of the theorem gives the corresponding result for 
higher derivatives. For instance, the series for/"(x) can be found by dif- 
ferentiating the series for f(x) twice, provided /"(x) satisfies the conditions 
of the theorem. 

We shall establish Theorem III by applying Theorem I to the function 
f(x). Being continuous, jf'(x) has a Fourier series, and the constant term 
a 0 can be found from 

TOO - f /'(*) dx « /Or) - /(-T) - 0. 

J — 7T 

Thus, the series for /'(x) has the form (25-4), namely, 

2(a n cos nx + b n sin nx). (25-10) 

It follows from Theorem I that the Fourier series for the function 

f f'(t)dt~f(x) -/(-*) 

J — T 

has the form (25-5), and hence the series for f(x) is 

/( — ir) + - Aq + 2) sin nx cos nxY (25-11) 

2 \n n / 

By inspection, we see that differentiating (25-11) gives (25-10). In other 
words, the Fourier series for f'(x) can be found by differentiating the series 
for /(x), and this is the main assertion in Theorem III. 

Since the differentiated series is a Fourier series, its convergence can be 
tested by the usual methods. In particular, if /'(x) satisfies the Dirichlet 
conditions and is continuous, then the Fourier series for f'(x) converges to 
/'(x). Thus, Theorem III is established. 

The foregoing methods lead to some important inequalities for the 
Fourier coefficients. When a function f(x) satisfies the Dirichlet condi- 
tions, it can be shown 2 that the Fourier coefficients have the order of mag- 
nitude \/n. That is, there is a constant M depending on f{x) but not on n 

1 It can be shown that if/'(x) satisfies the conditions of Dirichlet, then f'{x) is neces- 
sarily continuous. This follows from Darboux’s theorem. See, for example, L. Brand, 
“Advanced Calculus,” p. 112, John Wiley k Sons, Inc., New York, 1955. 

*§ee I. 8. Sokolnikoff, “Advanced Calculus,” p. 406, McGraw-Hill Book Company, 
Inc., New York, 1939. Cf, also Prob. 4. 



sbc. 25 ] 
such that 


ADDITIONAL TOPICS IN FOURIER SERIES 


211 


\a n 



\K\ 



( 25 - 12 ) 


Now, if the Fourier coefficients of f'(r) in (25-10) satisfy these conditions, 
then (25-11) shows that the coefficients of f(x) are bounded by M/ri‘ 
More generally, we can start with f (k) (x) and integrate k times. The con- 
stants of integration drop out as in the derivation of (25-10), and we 
obtain : 

Theorem IV. Let f(x) have period 2 tt and suppose the kth derivative of 
f(x) satisfies the conditions of Dirichlet on [ — ir,ir]. Then the Fourier coeffi- 
cients of f{x) satisfy the inequalities 


M 


\b n \ < 


M 

IkTi' 


where the constant M depends on f(x) but not on n. 


PROBLEMS 


1. By integrating the series (25-8) from 0 to x deduce that 

sin nx 


r{x 2 ■— tc~) « 12 £ (-I)*- 


»«*i n° 

2. By integrating the series m Prob. 1 from to x deduce that 


1 

48 


( T 2 - * 2 ) 2 


90 


£(-!)“- 


3 . Show that the following is not a Fourier series: 

^ sin nx 
n w 1 log (1 + n) 

Hint If it is a Fourier series, the integrated series must converge for all x. 

4 . Deduce (25-12) when /Or) is piecewise smooth on [ — Hint: Let the points 
Xk divide [ — ;r,7r] into a finite number of intervals on each of which /'(x) is continuous. 
The Fourier coefficients are obtained by adding integrals of the type 

f k J> 1 f(r) cos nx dx or f +1 /W sin nx dx t 

Jx k Jx k 

and these can be integrated by parts. 




CHAPTER 3 

FUNCTIONS OF SEVERAL VARIABLES 




The Technique of Differentiation 

1. Basic Notions 217 

2 Partial Derivatives 219 

3 Total Differentials 223 

4. Chain Rule 228 

5. Differentiation of Composite and Implicit Functions 230 

6. Higher Derivatives of Implicit Functions 235 

7. Change of Variables 237 

Applications of Differentiation 

8. Directional Derivatives 243 

9. Maxima and Minima of Functions of Several Variables 246 

10 Constrained Maxima and Minima 249 

11. Lagrange Multipliers 254 

12. Taylor’s Formula for Functions of Several Variables 257 

Integrals with Several Variables 

13. Differentiation under the Integral Sign 261 

14. The Calculus of Variations 264 

15. Variational Jboblems with Constraints 269 

10. Change of Variables in Multiple Integrals 270 

17. Surface Integrals 277 


215 




The considerations of the preceding chapters were confined primarily 
to functions y = f{x) of a single independent variable x. One does not 
have to go far to encounter functional relationships depending on two or 
more independent variables. In courses in analytic geometry and calculus 
the reader has learned that a functional relationship of the form z = f(x,y) 
may be represented as a surface, and lie has made use of partial derivatives 
to study some properties of surfaces. In this chapter the familiar concepts 
underlying the study of real functions of two variables are sharpened and 
extended to functions of many variables The bearing of such extensions 
on the calculation of rates of change and maximum and minimum values 
of functions of several variables is indicated in numerous problems of 
practical interest. 

The concluding sections of the chapter deal with integrals of functions 
of several variables They contain an introduction to the calculus of 
variations - a subject of great importance in physics and technology. 
Many situations can be characterized by statements to the effect that 
certain integrals attain extreme values. The determination of such ex- 
tremes is in the province of calculus of variations. 


THE TECHNIQUE OF DIFFERENTIATION 

1. Basic Notions. Let z ~ f{x,y) be a real-valued function of two inde- 
pendent variables (x,y). We can think of (x,y) as the coordinates of a 
point in the xy plane and interpret z as the height of the surface defined 
by z = f(x,y). The function f(x,y) may be determined for every point 
(; x,y ) in the xy plane, or the points for which it is determined may occupy 
a certain region R in that plane. 

For example, 

z — x 2 y 2 (1-1) 

represents the paraboloid of revolution for every pair of values (x,y) } while 

Z = (1-2) 

217 



218 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 


represents the surface of the hemisphere only for those values of (x,y) 
for which x 2 -f y 2 < 1. In the example (1-1) the region R is the entire 
xy plane, while in (1-2) it is the interior and the boundary of the unit 
circle x 2 + y 2 » 1 . The function 


1 


(1-3) 


is defined in the circular region x 2 + y 2 < 1, but not on the boundary 
x 2 + y 2 = 1. If the region of definition of the function includes its 
boundary C, we sliall say that the function is defined in the closed region R . 
When the boundary of the region R is not included, the region is said to 
be open . 

To define the continuity of z ~ f(x,y) at a given point we need the 
notion of the neighborhood of that point. The neighborhood of P(x Q ,y 0 ) 
is the set of all points P(x,y) interior to a circle with center at (x 0) y 0 ). 
If the radius of this circle is 8 > 0, then the neighborhood of (xodA>) is 
a circular region (.r — x 0 ) 2 + (y — ijo) 2 < 6 2 . The positive number 5 
can be chosen arbitrarily small. The extension of this definition to spares 
of more than two dimensions is immediate. The neighborhood of 
P(xo,yoZo) is the open spherical region 


(x - x 0 ) 2 + (y - ?/o) 2 + (z - zo) 2 < 5 2 . 


The neighborhood of the point Po{xo f yo } ZoJo) in the space of four variables 
x t y t z,t is the set of “points” (x } y,z>t) such that 

( x — To) 2 + (y — y«) 2 + (z — 2 0 ) 2 + (t — to) 2 < $ 2 , 

and so on for spaces of higher dimensionality. 

Intuitively the notion of continuity of z = /Or,y) at a given point 
PoixoiVo) means that the value of f{x,y) throughout a neighborhood of 
(xo t yo) will differ from f(x 0t y 0 ) by as little as desired if the neighborhood is 
chosen sufficiently small. In symbols this means that if one specifies a 
positive number e, no matter how small, then for all points in a certain 
circular region (x — x 0 ) 2 + (y — Vo) 2 < we have 

I f(x,y) - f(x 0 ,y 0 ) | < €. (1-4) 

An alternative notation for (1-4) is 

lim f(x,y) = f(x 0 ,y 0 ), (1-5) 

(x,v) — (ZqI/q) 

which states that as the point (x,y) is made to approach (x 0 ,y 0 ), the value 
of the limit is equal to the value of the function at (xo,j/o)- 
We extend this definition (1-5) to functions f(x u x 2 , . . .,x„) of n variables 
in the obvious way: A function /(x,,x 2) - • -,x») is continuous at the point 



219 


SEC. 2] THE TECHNIQUE OP DIFFERENTIATION 

P 0 (xi,X2,.. .,x”) whenever 

lim f(x u r 2 ,...,x n ) = /(*?,*?,.. .,x° n ). 

The “point P” here means the set of n real numbers (x!,x 2 , . . . ,x n ). Clearly, 
/(: r 1 ,^ 2 , * • * ,* r n) cannot be continuous at (x 1 ,x 2> • . - ,x n ) if it is not defined at 
that point. 

Whenever /(x lr r 2 ,. . .,j n ) is continuous at every point P of the given 
region /?, it is said to be continuous in the region R. Functions with which 
we shall deal for the most part will be continuous in some region, open or 
closed. 

PROBLEM 

Describe the regions of definition and the surfaces defined by the following functions z: 

(a) x — y 4- 2 * 1; (b) z » y; 

(r) y 2 + z 2 * 25; (d) z « 1 /(.r 2 + y 2 ) ; 

(e) z * \/x\ (/) z « Vi - (x - l) 2 - ?/ 2 . 

2. Partial Derivatives. Let u — f(x,y) be a function of two independent 
variables x, y , and let it be defined at a point (x<,,i/o) and for all values of 
(x t y) in some neighborhood of (x 0 ,t/o). If f/ is set equal to y<), then u be- 
comes a function of one variable x, namely, 

u = /(x,y 0 )* 

If this function has a derivative with respect to x, the derivative is 
called the partial derivative of f(x,y) with respect to x for y « ?/ n . In like 
manner, if x is assigned a constant value x.), the derivative with respect 
to y of the resulting function f(x () ,y) is called the partial derivative of f(x,y) 
with respect to y for x = x 0 . The customary notations for the partial 
derivative of u = /(x,y) with respect to x arc 

du df 

* 'U'Xf fxy and * 
dx dx 

The partial derivatives of a function /(x i,x 2 ,...,x n ) of n independent 
variables are obtained by fixing in it the values of n - 1 variables and 
computing the derivative of the resulting function of a single variable. 
Thus, 

SO,y) = yx* - 2 yx (2-1) 


has the partial derivatives 



220 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 


If « « f(x,y) is a function of two independent variables, it is easy to 
provide a simple geometric interpretation of partial derivatives u x and u v . 
The equation u — f(x,y) is the equation of a surface (see Fig. 1). If x 



Fio. 1 


is given a fixed value x 0 , u — f(x 0l y) is the equation of the curve AB on 
the surface formed by the intersection of the surface and the plane x — .r 0 . 
Then 

u = = j fj/ o, Vn + Ay) ~ /(t q,.Vo) 

V dy Ay -» o Ay 

is the slope at any point of AB. Similarly, if y is assigned the constant 
value y Qt then u « /(x,?/o) is the equation of the curve CD on the surface 
and 

du f (* o + At, ?y«) - f(io,y 0 ) 

dx ax -» o Ax 


is the slope at any point of CD. 

In Chap. 5 we shall see that the partial derivatives u x> u u> u x of u = 
f(x,y,z) can be interpreted as rectangular components of a certain vector, 
called the gradient of u. This vector provides a measure of the space 
rate of change of u. 

The partial derivatives f Xv f Xv . . f z% of f(x u x 2y ,. .,x n ) are functions 
of X 1; x 2 , . * x n; and they may have partial derivatives with respect to 
some or all of these variables. These derivatives are called second partial 
derivatives of /(x i,x 2 , . . . ,x n ). If there are only two independent variables, 
f(x,y) may have the second partial derivatives 


dx \dx) dx 2 

dx \dy/ dx dy 




THE TECHNIQUE OF DIFFERENTIATION 


221 


BEC. 2) 

It should be noted that f xv means that df/dx is first found and then 
d/dy(df/d x ) is determined, so that the subscripts indicate the order in 
which the derivatives are computed. In 

dy dz dy \dx/ 

the order is in keeping with the meaning of the symbol, so that the order 
appears as the reverse of the order iti which Ihe derivatives are taken. 

For the function f(x,y) in (2-1) we get, on noting (2-2), 

d / df\ d 

f*y ~ ~ ” ( " ~) ~ ™ (2*r// — 2 y) ~ 2x — 2, 
dy \dx/ dy 

d / a f\ a n 

fyr =—(--) = — (a- 2 - 2x) = 2r - 2 
dx \dy/ dx 

fyy ” ~~ (x* 2.r) * 0, 

dy 

d 

fxx - 7- (2 ry ~ 2 y) = 2 y. 

dx 

In this example f xy — f yt , and indeed, one rarely meets functions for which 
the so-called mixed derivatives are unequal. In fact one can prove i that 

ff ~ 

dx dy dy dx 

whenever these derivatives are continuous at the point in question. 

The process of defining partial derivatives of higher orders is obvious 
from the foregoing, and it is possible to establish equalities such as f xyx 
- /try ” fysx and f yxy = f xyv « fyy, whenever these derivatives are con- 
tinuous at the point in question. 

We note in conclusion that although the notation du/dx u>r the partial 
derivative v x suggests a quotient of some quantities analogous to f lie dif- 
ferentials dy and dx in the notation dy /dx for the derivative of y = f(x), 
no such interpretation is available for partial derivatives. To stress the 
point that du/dx should never be thought of as a fraction, we give an 
example. 

E ram pie. Consider the equation for an ideal gas pv * RT y whole p is the pressure, 
v is the volume, T is the absolute temperature, and A' is a physical constant. It should 
be noted first that the concept of partial derivatives hinges on the agi cement as to which 
variables in a given functional relationship are assumed to be independent. Thus, if 

‘See I. S. Sokolnikoff, “Advanced Calculus/' sec. 31, McGraw-Hill Book Company, 
Inc., New York, 1939. 



222 FUNCTIONS OF SEVERAL VARIABLES 

we solve our gas equation for p, we obtain 


[chap* 3 


We can then compute 


dp RT 

dv " v 2 


dp R 
~dT v 


On the other hand, if we solve for v t we get 


in which p and T are now regarded as the independent variables, and we can, therefore, 
COmpUte dv R dv RT 

ar “ p’ ap“ p 2 ' 2-4 

We can also solve for T and get 


in which p and v are to be considered as the independent variables, so that 

dT v dT p 

dp R dv R K ; 

From Eqs. (2-3) to (2-5) we obtain 

dp dv HZ _ ZL ?L v 

dv dT dp 1 ? p R ' 

since pv m RT. But if it were possible to treat the terms in the left-hand member of 
(2-6) as fractions, we should have obtained +1. 


PROBLEMS 

1. Find dz/dx and dz/dy for each of the following functions* 

(a) z « y/x; (b) z ** x?y + tan -1 (y/x); ( c ) z - sin xy 4~ j; (d) z « e x log y; 
(e) z x 2 y 4“ sin" 1 x 

2 . Find dit/dx , du/dy, and du/dz for each of the following functions: 

(a) u *» x 2 y 4 - yz — xz 2 ; (b) u *» xyz + log xy; 

(c) u *■ z sin" 1 (xfy); ( d ) u « Or 2 4- y 2 4- z 2 )^; 

(e) u * (x 2 4- y 2 + z 2 )~ h . 

8. Verify that d 2 f/dx dy *»* d 2 fjdy dx for 

(a) / ** cos xy 2 , (b) f » sin 2 x cos y , (c) / =* e v/x . 

4 . Prove that if 

(a) f(x f y) « log (x 2 4- y 2 ) + tan" 1 -» then 4- = 0; 

x dx* dy 2 

W f(x,V,*) - (x 2 4- V 2 + z 2 )“~ H , then 4* ~ 0. 

dx 2 dy 2 dr 



SEC. 3] THE TECHNIQUE OF DIFFERENTIATION 223 

3. Total Differentials. The differential dy of a function y = f(x) is 
defined by the formula 

dy = f'(x) dx, (3-1) 


where dx s Ax is an arbitrary increment of the independent variable x. 
We agree to call an increment of the independent variable x the differential 
of x. 


Since (Fig. 2) 


/(i+Ai) 

fix) 

JP. 


f f Ax 

JzJT 


\ 

dy 


| x x+Ax x 


Fig 2 


, v /(x + Ax) — f(x) 

f (x) ~ hm ~ - hm ~ — — 

Ax — * 0 Ax Ax — * 0 Ax 


(3-2) 


we can write, on recalling the definition of the limit, 


Ay 

v = /'<*> + «> 

Ax 


(3-3) 


where lirn « — 0. Hence 

Ax 0 

Ay ~ /'(x) Ax + t Ax. 
The substitution from (3-1) in (3-4) then yields 

A?/ - dy + € Ax, 
lim € = 0 as Ax — » 0 . 


(3-4) 


(3-5) 


Figure 2 illustrates geometrically the relations between Ay, dy, and dx, 
and formula (3-5) shows that for small values of Ax, the increment Ay 
is a good approximation to the differential dy in the sense that 


Ay - dy 


Ax 


(3-6) 


where e — ► 0 as Ax — ► 0. 

One can construct a similar approximation to the increment A u for 
the function u ~ /(x,y) when x and y are allowed to acquire the respective 
increments Ax and Ay. 



224 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 

The presentation of the essential ideas in this construction is greatly 
simplified by the use of the mean-value theorem of the differential calculus. 
This theorem states that whenever fix) has the derivative fix) at every 
point of the interval (x, x + Ax), then 

fix + Ax) — f{x) 

1 — 7 — — - m (3-7) 

Ax 

where £ is an intermediate point in the interval. The geometric meaning 
of this theorem is exceedingly simple. Formula (3-7) states that the slo|>e 

fjx + Ax) - fjx) 

Ax 

of the secant line AB (Fig. 3) is equal to the slope /'(£) of the tangent 



line CD to the curve y = fix) at an intermediate point £ in the interval. 
Since £ =* x + 0 Ax, where 0 Ax is some fraction of the length Ax, we 
can write (3-7) as 

fix + Ax) - fix) « /'(x + 0 Ax) Ax, 0 < 0 < 1. (3-8) 

Consider now a function u = f(x } y) of two variables. The increment 
A u that results from replacing x by x + Ax and y by y + Ay is 

A u fix + Ax, y + Ay) - f{x,y), (3-9) 

If we add and subtract fix, y + Ay) in the right-hand member of (3-9), 
we obtain 

Aw ~ [/(x + Ax, y + Ay) - /(x, y + Ay)] + [fix, y + Ay) ~~ fx,y)l 

(3-10) 



THE TECHNIQUE OF DIFFERENTIATION 


SEC* 3] THE TECHNIQUE OF DIFFERENTIATION 225 

The expression in the first pair of brackets in (3-10) is the increment in 
the function f(x, y) when the second variable in it has a fixed value y + Ay, 
Accordingly, we can apply formula (3-8) to it and write 

f(x + A x,y+ Ay) - f(x, y + Ay) = f x (x + 0i Ax, y + Ay) Ax, (3-11) 

where 0 < 0* < 1. 

Similarly, the application of (3-8) to the expression in the second set 
of brackets in (3-10), in which x has a fixed value, yields 


f(x, y + Ay) - f(x,y) = f v (x, y + 0 2 Ay) Ay, 


(3-12) 


where 0 < d 2 < 1. 

Now if the partial derivatives / X (x,y) and/ v (ay/) are continuous functions, 
then 

f z (x + 0iAx,y± Ay) = fjx,y) + e h 

, (3-13) 

/*(■*, y + 02 Ay) « f v U,y) + « 2 , 

where lim = 0 and Jim « 2 = 0 as Ax and Ay approach zero. Hence we 
can write (3-11) and (3-12) in the forms 

}{x + Ax, y -f Ay) - /(x, y + Ay) = \f x {x,y) + €l ] Ax, 

/(x, y + Ay) - /(x,y) * [f v (x,y) + t 2 ] Ay, 

so that (3-10) becomes 

Aw = f x (x,y) Ax + f v (x,y) Ay + t x Ax + e 2 Ay. (3-14) 

If we define the differential du of u = /(x,y) by the formula 

du - f x Ax + f v Ay (3-15) 

df df 

as — Ax H Ay 

dx dy 

we can write (3-14) in a form analogous to (3-5): 

A u — du + €i Ax + *2 Ay, 

lim e x = 0, lim e 2 — 0 as Ax —+ 0 and Ay 0. 

Formula (3-10) shows that when the increments Ax and Ay are small, the 
differential du is a good approximation to A u in the sense that 


(3-16) 


Aw — du 


ci Ax + e 2 A y 


Vl&z? + (&V) 2 V (Ax ) 2 +W 2 


o 


as Ax and Ay approach zero. 

As in the case of functions of one independent variable, we agree to 



226 FUNCTION’S OF SEVERAL VARIABLES [CHAP. 3 

write the increments Ax and Ay in the independent variables as dx and dy, 
respectively. Then (3-15) reads 

df df 

du ~ — dx + — dy, (3-17) 

dx dy 

Whenever (3-16) holds, the function u = f(x;y) is said to be differentiable , 
and du in (3-17) is called the total differential 1 A function which is dif- 
ferentiable at each point of a region is said to be differentiable in the region . 
The foregoing discussion shows that a function f(x,y) is differentiable 
whenever the partial derivatives f x and f v are continuous. 

The foregoing considerations can be extended to functions u = f(x i,x 2 , 

. . . t x n ) of n independent variables. The total differential du is given 
by the formula 

df df df 

du — — dx i H dx 2 H 1 dx n (3-18) 

dx i dx 2 dx n 

whenever the partial derivatives f Xx are continuous functions. 

It should be noted that the total differential du is equal to the sum of 
n terms involving independent increments dx t When a number of small 
changes are taking place simultaneously in a system, each one proceeds 
as if it were independent of the others, and the total change is the sum of 
the effects due to the independent changes. Physically, this corresponds 
to the principle of superposition of effects. 

Example 1. Find the total differential of u « e s yz 2 . Since u Xt Uy , u x are obviously 
continuous functions, formula (3-18) yields 

du « e*yz L dx -f e 1 * 2 dy F 2e J yz dz. 

Example 2. A metal box without a top has inside dimensions 6 by 4 by 2 ft. If the 
metal is 0.1 ft thick, find the actual volume of the metal used and compare it with the 
approximate volume found by using the differential. 

The actual volume is A V , where 

AF * 0.2 X 4.2 X 2.1 - 6 X 4 X 2 - 54.084 ~ 48 « 6.684 ft*. 

Since V ** xyz t w here x « 6, y » 4, z *» 2, 

dV ■* yz dx -f xz dy -f xy dz 

* 8(0.2) -f- 12(0.2) -f 24(0.1) « 6.4 ft*. 

1 In Chap. 5 we shall encounter expressions of the form (3-17) in which f x and f y be- 
come discontinuous at certain points of the region and u is a multiple- valued function. 
Such expressions are generally called exact differentials, and they are also denoted by the 
same symbol du. For technical reasons, explained in Chap. 5, it is usually necessary to 
assume the continuity of f x and f y , in which event the terms exact and total differentials 
become synonymous. A geometric meaning of the differential (3-17) is given in Sec. 
10, Chap. 4. 



THE TECHNIQUE OF DIFFERENTIATION 


SEC. 3] 


227 


Example 3. Two sides of a triangular piece of land (Fig. 4) are measured as 100 and 
125 ft, and the included angle is measured as 60°. If the possible errors are 0.2 ft in 
measuring the sides and 1 0 in measuring the angle, what is the approximate error in the 
area? 



Fro. 4 

Since A * Yixy sin a, 

dA =* l / 2 {y sin a dx -f x sin a dy -f xy cos a da) t 
and the approximate error is therefore 

/T A 

dA ~l \ 125 (- 2 “) (°-2) + 100 (' 2 ~~) (°-2) + 100(125) Q T ^] - 74.0 ft*. 


PROBLEMS 

1. A dosed cylindrical tank is 4 ft high and 2 ft in diameter (inside dimensions). 
What is the approximate amount of metal in the wall and the ends of the tank if they 
are 0.2 in. thick? 

2 . The angle of elevation of the top of a tower is found to be 30 °, with a possible error 
of 0.5°. The distance to the base of the tower is found to be 1,000 ft, with a possible 
error of 0.1 ft. What is the possible error in the height of the tower as computed from 
these measurements? 

3 . What is the possible error in the length of the hyjiotonuse of a right triangle if the 
legs are found to be 11 5 and 7.8 ft, with a possible error of 0.1 ft in each measurement? 

4. The constant C m Boyle’s law pv = C is calculated from the measurements of p 
and v. If p is found to be 5,000 ib per ft 2 with a possible error of 1 per cent and v is 
found to be 15 ft 3 with a possible error of 2 per cent, find the approximate possibleerror in 
C computed from these measurements. 

6. The volume v , pressure p, and absolute temperature T of a perfect gas are con- 
nected by the formula pv * R7\ where R is a constant. If 7" » 500°, p «= 4,000 lb per 
ft 2 , and v *» 15 2 ft 3 , find the approximate change in p when T changes to 503° and v 
to 15.25 ft 3 . 

6. In estimating the cost of a pile of bricks measured as 0 by 50 by 4 ft, the tape is 
stretched 1 per cent beyond the estimated length If the count is 12 bricks to 1 ft 8 and 
bricks cost $20 per thousand, find the approximate error in cost. 

7 . In determining specific gravity by the formula ,«? = A /(A — W) f where A is the 
weight in air and W is the weight in water, A can be read within 0.01 lb and IF within 
0,02 lb. Find approximately the maximum error in s if the readings are A » 1.1 lb 
and W ** 0.6 lb. Find the maximum relative error As/s. 

8. The equation of a perfect gas is pv » RT. At a certain instant a given amount of 
gas has a volume of 16 ft 3 and is under a pressure of 36 psi, Assuming R * 10.71, find 
the temperature T. If the volume is increasing at the rate of % efs and the pressure is 
decreasing at the rate H psi per sec, find the rate at which the temperature is changing. 



228 FUNCTIONS OF SEVERAL VARIABLES 

9. The period of a simple pendulum with small oscillations is 


[chap. 3 


If T is computed using l 8 ft and g — 32 ft per sec per sec, find the approximate 
error in T if the true values are l * 8.05 ft and g ** 32.01 ft per sec per sec. Find also 
the percentage error. 

3.0. The diameter and altitude of a can in the shape of a right circular cylinder are 
measured as 4 and 6 m., respectively. The possible error in each measurement is 0.1 in. 
Find approximately the maximum possible error in the values computed for the volume 
and the lateral surface. 

11. We define an approximate relative error ( m the differentiable function / by the 
formula e ** df If. Show that the approximate relative error of the product is equal to 
the sum of the approximate relative errors of the factors. Hint: e * d log/. 

4. Chain Rule. Let w = f(x,y) be a function of the variables x and y 
which, in turn, are functions of some independent variable If t is given 
an increment A f, the functions x and y will acquire increments Ax and Ay> 
and consequently u will receive an increment Au. 

Assuming that u ~ f(x,y) is continuous together with its partial deriva- 
tives, one can write [see (3-14)] 


du du 

Au = — Ax H Ay 4- Ax + *2 A?/. 

dx dy 

Dividing both sides of this expression by At gives 

Au du Ax du Ay Ax Ay 

as 1 b «1 * 4“ *2 * 

At dx At dy At At At 


(4-1) 


Now if it is supposed that x and y can be differentiated with respect to l, 
the expression (4-1) gives, upon passing to the limit as At —> 0, 

du du dx du dy 
dt dx dt dy dt 


df dx df dy 
dx dt dy dt 


since ~ » 0 and e 2 — ► 0. The reason for the vanishing of €j and « 2 as 
At 0 is as follow’s. Since x and y are assumed to be differentiable func- 
tions of t , the identities 


Ax Ay 

Ax *» — At, Ay * — At, 

At At 

show that Ax ~ * 0 and Ay — * 0 as At — ► 0. Rut when A# — ► 0 and 
Ay 0 we know that n and <? 2 approach zero by (3-16). 



THE TECHNIQUE OF DIFFERENTIATION 


229 


SEC. 4] 

Formula (4-2) gives the rule for the differentiation of composite functions. 
It is clear that if u is a function of a set of variables, x x , a* 2 , . . . , x n , where 
each variable is a differentiable function of an independent variable t 1 
the derivative of u with respect to t is given by the chain rule: 

du du dx i du dx 2 Ou dx n 

— 1 1 1 

dt dxi dt dx 2 dt dx n dt 

A special case of formula (4-2) is of interest. If it is assumed that 
t = O', (4-2) becomes 

du du du du 

T = -T + T ~T ' ( 4_4 ) 

dx dx dy dx 

Formula (4-4) can be used to calculate the derivative of the implicit 
function given by 

fU*V) - 0. (4-5) 

Let it be assumed that (4-5) can be solved tor y to yield a real solution 

y = <p(x); (4-6) 

then the substitution of (4-6) in the left-hand member of (4-5) gives an 
identity 

0 = f(x t y) 9 where y = <p{x). (4-7) 


Applying (4-4) to f 1-7) gives 


and solving for dy/dx , 


df df dy 
0 - - -- 
dx dy dx 


The formula (4-8) assumes that df/ dy does not vanish for the point 
(x 0) yo) at which the derivative is calculated. 

Example 1. ljetf{x,y) « 3jr 8 y 2 4- * cos y *» 0; then 


9 x l y l 4* cos y, 


i\x*y — x sin y 


dy 9x~id -f cos y 

so that r- ** •" ;r -f 

dx bx 3 y - x sin y 

for all values of x and y that satisfy the equation 

3 x^y 2 4- x cos y * 0 

and for which 6 x z y ~~ x siri y 0 . 

Example 2. Let x 2 4- y 2 - 0; then df/dx - 2x , df/dy - 2 y. But it does not follow 
that 

dy x 



230 


FUNCTIONS OF SEVERAL VARIABLES 


{CHAP, 3 

This result is absurd inasmuch as the only real values of x and y that satisfy x 2 •+- y* ■* 0 
are x ** 0 and y » 0. Since df/dy vanishes at this point, the formal procedure used in 
obtaining dy/dx m meaningless. 

Example 3, Let f(x } y) * 0 represent the locus of a curve, and let P(xo,yo) be a point 
on the curve. The equation of the tangent line to the curve at the point P is 


V - l/o 



(x - a?o), 

x-xo 


It follows from (4-8) that this equation can be written in the form 


fx(x o,yo)(x - xq) +f v (xo t y Q )(y - y 0 ) * 0. 


PROBLEMS 


1- Find the equation of the tangent line to the ellipse 

at the point (xo,s/o). 


x y M 
a 2 ^ l > 2 


1 


2 . Find the equation of the tangent line to the folium of Descartes 

x 3 -f V s — 3a xy ** 0. 


Note particularly the behavior of the tangent line to the folium at (0,0). 

3 . Find du/di if 

\x » e* + 


tan“ 


and 


4 . Find the equation of the tangent line to the ellipse 


x =* a cos 0, 
y ** b bin 0, 

at the point where 0 * x/4. 

6. (a) Find du/dt, if u = e 1 sin ?/£ and x » f 2 , y « £ — 1, z ® 1 //; 

(6) find du/dr and du/dB , if a =* x 2 — 4i/ 2 , x = r sec 0, and y = r tan 0. 

6. (a) Find du/dx and du/dx , if u « x 2 -f y 2 and y * tan x; 

(6) given F ** f(x,y,z), where x *» r cos 0, i/ ** r sin 0, z ** t; compute dV/dr , dF/d0 f 
dV /dt in terms of dF/dx, dV/dy f and dF/dz. 


6. Differentiation of Composite and Implicit Functions. The reasoning 
employed in the preceding section can he applied in obtaining the total 
differential, and hence the derivative, of a function of n variables 


u =/(*i, x 2 ,...,x n ), 

where Xi =* Xi(t), i ~ 1, 2, . . n 

are n differentiable functions of a single variable t. The resulting expression 
for the total differential is 



SEC* 5] THE TECHNIQUE OF DIFFERENTIATION 231 

A question arises concerning the validity of formula (5-1) in the case 
where the variables x t are functions of several independent variables 
*!, < 2 , t m . Thus, let 

u « f(x i,x 2 ,...,x n ) (5-2) 

be a function of the n variables r t , where the x t are functions of the vari- 
ables t l} t 2 , . tm, say 

X t = X t (ti f l 2} , » -fm)) i ^ 1,2, . . » , 71. (5-3) 

If all the variables except one, say t kl are held fast, (5-2) becomes a function 
of the single variable t k and one can calculate the derivative df/dt k with 
the aid of (5-1). The notation df/dtk , instead of df/dt k , is used to signify 
the fact that all variables except t k are held fast 
Assuming the continuity of the derivatives involved, one can write 


df 

df dx, 

, df dx 2 


df dx n 


see — 

H — + • 




at i 

dxi dt\ 

dx 2 dt 1 

dx n dt\ 

df 

df dx i 

, df dx 2 


df dx n 


zrx — 

+ — h • ■ 



dt 2 

dx ! dl 2 

dx 2 dt 2 

dx n dt 2 


df df dxi df dx 2 df dx n 

dim dxi dt m dx 2 dt m dx n dt m 

If df j dt , , df/dt 2 , df ' dt m are multiplied, respectively, by dt\, di 2l 

. . . , dt m and the resulting expressions added, one obtains 


dli 


+ 

+ 

+ 


df 


d f , 



dt 2 + ■ ■ ■ 

H dt m 


dt 2 


dim 


df 

(dX\ 

dli 

dxi 


( - dli 

+ - - dt 2 + • 

■ * H 1 

dXj 

\dt\ 

dl 2 

St m 

df 

(dx 2 

, dx 2 , , 

dx 2 


(- -dti 

_j -j- • • 

■ -i 1 

Ox 2 

\dt i 

dt 2 

dl m 

df 

(dx n 

dx n 

dx n 


( — dt , 

-j dt 2 + • 


dX n 

\di, 

dt 2 



dtmj 

dt^\ 


dtn 


The left-hand member of this expression is the total differential of f{x\ } x 2} 
. . .,x n ), regarded as a function of the independent variables t x , t 2 , . . t mt 
whereas the terms in the parentheses in the right-hand member are pre- 
cisely the total differentials of (5-3). Hence, one can write 

df df df 

df « — dx i H dx 2 + * • • + ” — dx n , 

dx i dx 2 dx n 



232 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 

which shows that formula (5-1) is valid whether the x t -s are the independent 
variables or are functions of any set of other independent variables. 

The foregoing can be summarized as follows: 

Theorem. If u « f(x 1; x 2 , . . . r x«), then 

df df df 

du = — dx i 4 dx 2 H i dx n , 

dxi dx 2 dx n 

regardless of whether the variables x,- are the independent variables or are 
functions of other independent variables f*. It is understood that all the 
derivatives involved (the df/dxk and dxjdtk) arc continuous functions. 

The fact that the total differential of a composite function has the same 
form irrespective of whether the variables involved are independent or 
not permits one to use the same formulas for calculating differentials as 
those established for the functions of a single variable. Thus, 

d(u + v) = du + dv } 

d(uv) d(uv) 

d(uv) = du H dv 

du dv 


and so forth. 


v du + u dv, 


Example 1. If u *» xy 4- yz -f zx, x ** l, y e *, and z » cos l, 

du (lx K dy dz 

^-( ! /+*)- + ( I +z ) - + (x + 3 , ) - 

“» («“* 4- cos 00) 4* (t + cos 0( —e" 1 ) 4- 0 4* c “■**)( — sin t) 

e~* 4* cos t - fe*~* cos t — t sin * — sin 

This example illustrates the fact that this method of computing du/dt is often shorter 
than the old method in which the values of r, y, and z in terms of t are substituted in 
the expression for u before the derivative is computed. 

Example 2. If f{x t y) *■ x 2 4“ V 2 , where x * r cos ? and y *= r sin ?>, then 

a/ df dx df dy „ . , 

— „ j » 2.r cos V s + 2y sin <p *■ 2r cos 2 y> 4- 2r sin 2 *» 2r. 

Br dx dr By Br 


df_ _ afax 

dip dx dtp dy dtp 


2x(—r sin <p) 4- 2 y{r cos ip) 


-2 r 2 cos sin 4* 2r 2 cos sin » 0. 
or df ** 2x dx 4- 2y dy. 


Also, df a 2 r dr or df » 2x dx 4- 2y dy. 

Since f{x,y) * x 1 4* V 2 ** r 2 , these results could have been obtained directly. 
Example 3. Ijet z » e**, where i *» log (u 4- ®) and y «* tan~ l ( u/v ). Then, 


cto w4» 


dy v 

du v 2 4 - u l 



THE TECHNIQUE OF DIFFERENTIATION 


233 


SEC. 5] 
Hence, 


Similarly, 


dz 

dll 


dz dx dz dy 
dx du dy du 


ye** xe^v 
u *+■ v v 1 + 


dz ye ** xc**u 

dv u -f v v 2 -f u 2 


The same results can be obtained by noting that 

dz «■ i/e Ty dx -f xe** cty. 

, dx dr 1 1 

But dx * — au H at? » aw 4- dv 

du dv u 4- v u 4- v 


and 

Hence, 


But 


dy « — du + Q dv 


du 


dv 


v 2 -i u‘‘ 


du - 


i 2 H u 2 


,dv. 


du -f ~ dv v du - // dv 

dz * ye* v f xr rv 2 

w -h f> tr 4" xr 


( ye * V X. xe * Vv \ a 4 ( ,,cXV X€ xu u \ 

\m 4* v if 1 4- uv \a 4 v v 2 4 w 2 / 


dv. 


_ a^ 

dz - an \ - dr, 

a?i <>c 


and since du and dv are independent differentials, equating the coefficients of du and 
dv m the two expressions foi dz gives 


dz 

y ( .rv 

- ■ T - 4 

u 4- »» 

xr TV v 

dw 

v 2 + 1/2 

dz 

ye 1 * 

xe TV u 

dv 

u 4 v 

v 2 4 ir 


Let f(x,y,z) — 0 define any one of the variables as an implicit function 
of the remaining ones. If x and y are thought to be the independent 
variables and one can obtain a rea* differentiable solution for z in terms 
of x and y y it is possible to write 

dz dz 

dz = — dx H dy. 

dx dy 


But 


df df df 

df « — dx H — dy 4 dz 

dx dy dz 


0 . 


Substituting the value of dz in this equation gives 



FUNCTIONS OF SEVERAL VARIABLES 


234 


[chap. 3 


Since x and y are independent variables, we get, on setting in turn dx 
and dy =* 0, 


df df dz 
— + - — = 0 
dx dz dx 


0 


and 


df df dz 
dy dz dy 


= 0 . 


These equations could have been obtained directly by applying the chain 
rule to the equation f(x,y,z ) = 0, in which z is regarded as a function of 
x and y, but we wished to illustrate another procedure followed in Sec. 10 
and elsewhere. If df/dz ^ 0, these equations give 


dz df/dx dz _ df/dy 

dx df/dz dy df/dz 


(6-4) 


The formulas (5-4) permit one to calculate the partial derivatives of 
the function z defined implicitly by an equation 

f(w) = 0. 

As an illustration, let 

x 2 + 2?/ - 3 xz +1=0. 

Then, by (5-4), 

dz 2x — 3 z dz Ay 

dx —3* dy —3x 


Example 4. A function f(xi,X 2 , • . .,Xn) of n variables ii, xj, . . x n is said to be homo- 
geneou ? of degree m if the function is multiplied by \ w when the arguments x h x 2 , 
x n are replaced by Xxi, Xx 2 , . . ., Xx n , respectively. For example, f(x } y) *» x 2 /Vx 2 ~f y* 
is homogeneous of degree 1, because the substitution of Xx for x and X y for y yields 
Xx 2 /Vx 2 -f V- Again, » 0 /y) -b (log x - log y)/x is homogeneous of degree 

—1, whereas f(x,y,z) » z 2 /\/ x 2 + y 2 is homogeneous of degree %. 

There is an important theorem, due to PJuIer, concerning homogeneous functions. 

Euler’s Theorem. If u ** fix^xi , . . .,x n ) homogeneous of degree m and has con- 
tinuous first partial derivatives , then 

df df df 

Zl h x 2 - 1 b Xn — * mf(xi,Z 2 , . . . ,x»). 

dXi vJT 2 TOfi 


The proof of the theorem follows at once upon substituting 

x[ *• Xxi, xi — Xx 2l . . . , ■* Xx n . 


Then, since /(xi,xj,. . .,x») is homogeneous of degree m, 

ffeittyit • • *■* \ m f{x i,Xj, . . . ,x n ). 



235 


BBC. 6] THE TECHNIQUE OP DIFFERENTIATION 

Differentiating with respect to X gives 

df df df 

i^ Xi + ^* 2+ "' + ^ 1 " - mX "“^ x *•*» ■*«>• 

If X is set equal to 1, then x\ ■» xx, xa *» xa, . . x n « and the theorem follows. 


PROBLEMS 


1. (a) Find dy/dx if x sec y -f x 3 y 2 «■ 1; 

(6) find d«/dx and dz/dy if x 3 y — sin 2 -f z 8 * 0 

2. If / i^a function of w and v, where u « Vx 2 -f y 2 and t> - tan~ A (y/x), find d//dx, 
af/ay, V(df/axf+~(df/dy) 2 . 

8. If / is a function of u and v , where u ** r cos a and f» « r sin 0, find 


f/ df t 

dr* 60* 



4. If x * x' cos 0 y 1 sin 6, y «* x' sin 9 -f j/' cos prove that 

( 3 ,+ ®‘*(£)‘+( 5 )‘ 

5. Find the total differential if a ■» x 2 -f- y 2 , x » r cos 6, and y » r sin 

6. If / « e* 1 ', where x * log (u 2 + ^) h and y * tan"* 1 (u/v), find df/du and df/dv. 

7. If z « (n -f e)/(l — we), u « y sin x, and r *» e v>r , find dz/dx and dz/dy . 

8. Find dz/dr and dz/ds if z ** (x — y)/(I + xy), x = tan (r — s), and y « c rt . 

9. Verify Euler’s theorem for each of the following functions: 


(a) f(x t y t z) 

(» /(*,y) - 

(c) /(x,y) - 

(d) f(x,y t z) 


» x 2 y -f xy 2 + 2xyz; 

V 


JL _l IgjLf ~~ log y 

y 2 x* 


Vx 2 - y 2 


(e) f(x,y,z) 
(/) /(x,y) - 

(y) f(x,y) * 
(*) f(x t y) = 


- (x 2 + y 2 T z 2 )~ H ; 
V jc -f y ^ 

y 

x 2 -f y 2 


6. Higher Derivatives of Implicit Functions. The problem of calculating 
the derivative of y with respect to x when y is an implicit function of the 
independent variable x defined by 

f(x,y) * 0 (6-1) 

was discussed in Sec. 4. It was shown there that 

dy 

fx(x y y) + fv(x;y) 7 = 0. (6-2) 

ax 

Differentiating this equation again and assuming that all the derivatives 
involved are continuous functions of x and y give 

dy / dy \ 2 d 2 y 

f,x(x,y) + 2 f xv {x,y) — +f vv (x,y) + f v (x,y) = 


0 . ( 6 - 3 ) 



236 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 

If fy(x,y) 0 at the point where the derivative is desired, (6-3) can 
be solved for d 2 y/dx 2 and the value of dy/dx substituted from (6-2). The 
result is 

A ^ _ fxxfl - VxJJy + fyj'i 

dx 2 fl 

The process can be continued to obtain the derivatives of higher orders. 

A similar procedure can be employed to calculate the partial derivatives 
of a function z of two independent variables x and y defined implicitly 
by an equation of the form 

f(x,y,z) = 0. (6-4) 

Differentiating (6-4) with respect to x and y in turn gives 


Mx,y,z) + f z (x,y,z) 


dz 

dx 


= 0, 


fv(*>y>z) + fz{*,y,z) 


dz 

dy 


o. 


(6-5) 


If f z (x y y,z) does not vanish for those values of x, y } and z that satisfy 
(6-4), then Eqs. (6-5) can be solved for dz/dr and dz/dy. Partial deriva- 
tives of higher order can then be obtained by differentiating equations 
(«). 


Example Let it be required to find the derivatives of second order of the function z 
defined implicitly by the equation 


x 2 y l z~ 

h — 4- - 

a 2 ^ b 2 ^ c 2 


1. 


Differentiating this equation with respect to x and y gives 


2x 2 z dz 
a 2 r* dx 


0, 


2 y 2 z dz 
¥ + r“* oi/ 


o. 


Differentiating the first of Eqs (0-6) with respect to x and y, one obtains 
2 2 /dz\ 2 2 zd‘‘z 


2 dz dz 2 z dh 

~g H — r - *» 0. 

c 2 dx dy r dx dy 

Solving for dh/dx 1 and dhfdx By and making use of (6-6), one obtains 

d l z e z art 1 -f c V 


m) 



BBC* 7] THE TECHNIQUE OP DIFFERENTIATION 237 

In a similar way the differentiation of the second of Eqs. (6-6) with respect to y yields 

d 2 z c 2 b 2 z 2 + c?y 2 

dy 2 i> 4 r 8 


PROBLEMS 


1. Find y\ y'\ y ,n if x* -f y* — 3axy « 0. 

2. Find dz/dx, dz/dy, d 7 z/dx 2 , d 2 z/dx dy, and d 2 z/dy 2 at (1,1,1) if x 2 — y 2 -f- « 2 « 1. 

3. Find dz/dx , if 

(a) xz 2 — i/z 2 d- xj/ 2 z — 5 «* 0; (6) xz 8 - i/z -f~ 3xj/ « 0. 


7. Change of Variables. The main purpose of this section is to develop 
manipulative skill in calculating the derivatives of implicit functions and 
to indicate the formal modes of attack on the problem. The continuity 
of the functions and their partial derivatives is assumed throughout this 
section and will not be referred to again. 

T 

w * f(u,v) (7-1) 


denote a function of two independent variables u and v, and suppose that 

u and v are connected with some other variables x and y by means of the 

relations , N 

X = x(?/,v), 


V = 2f(w,»). 


(7-2) 


If Eqs. (7-2) are solved for x and y to yield 

u = u(x,y), 
v = v{x,y) t 


(7-3) 


and the expressions (7-3) are substituted for a and v in (7-1), there will 
result a function of x and y r say 


w = F(x } y). 


(7-4) 


The partial derivatives of w with respect to x and y can be calculated 
from (7-4) directly, but frequently it is impracticable to obtain the solu- 
tion (7-3), and we consider an indirect mode of calculation. By the rule 
for the differentiation of composite functions, 


dw 

du 

dw 

dv 


dw dx dw dy 

dx du dy du 

dw Ox dw dy 

dx dv dy dv 


(7-5) 


The partial derivatives dx/du, dy/du, dx/dv, and dy/ dv can be calculated 
from (7-2), and hence they may be regarded as known functions of u and 



238 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 


v . The partial derivatives in the left-hand members of (7-5) are also 
known functions of u and v, since they can be calculated from (7-1). 

Hence, equations (7-5) may be regarded as linear equations for the 
determination of dw/dx and dw/dy. Assuming that the Jacobian J(u f v) 
defined by 

\dx dy\ 


J(u,v) m 


du du 
dx dy 


dv dv 


is not zero and solving by Cramer’s rule give 



dw dy 


dx dw 


du du 


du du 


dw dy 


dx dw 

dw 

dv dv 

dw 

! dv dv 

dx 

J(u,v) 

dy 

J (u,v) 


The resulting expressions for dw/dx and dw/dy are known functions of 
u and v and thus can be treated exactly like (7-1) if it is desirable to cal- 
culate the derivatives of higher orders. 

As an example, consider the function w(r,6), and let it be required to 
calculate the partial derivatives of w with respect to x and ?/, where x — 
r cos 0 and y ~ r sin 8. Now 


dw 

dw 

dx 


dw 

dy 

dw 

dw 


— = 

= — 

— 

+ 

— 


= — cos 8 + 

— sin 0, 

dr 

dx 

dr 

dy 

dr ~ 

dx 



dw 

dw 

dx 


dw 

dy 

dw 


dw 

— = — 

— 

+ 

— 


= rsin 

o + 

— r 

be 

dx 

d6 


dy 

dd ~ 

dx 

dy 


Solving these equations for dw/dx and dw/dy in terms of dw/dr and dw/dd 
gives 


dw dw sin 6 dw 

— = cos 8 * 

dx dr r d8 

dw dw cos 8 dw 

— = sin 8 1 

dy dr r d$ 

The Jacobian J is, in this case, 

cos 8 sin 8 

— rsind rcos0 



which does not vanish unless r = 0. 



SEC. 7] THE TECHNIQUE OP DIFFERENTIATION 239 

As a somewhat more complicated instance of implicit differentiation, 
consider a pair of equations 


F(?,y,u,v) * 0, 
G(x,y y u,v) = 0, 


(7-6) 


and let it be supposed that they can be solved for u and v in terms of x 
land y to yield 

u = u(x,y), 

(7-7) 

v = v(x f y). 


The partial derivatives of u and v with respect to x and y can be obtained 
in the following manner. Considering x and y as the independent \ariables 
and differentiating Eqs. (7-6) with respect to r and y give 


dF 


dF du dF dv 


dF 

OF du 

dF dv 

— 

+ 

— 

= o, 

— + 

b 

= 0 , 

dx 


dv dx dv dx 

dy 

du dy 

dv dy 

dG 


dG du dG dv 


dG 

dG du 

dG dv 

— 

+ 

— ~ 

= 0 , 

— + 

j_ 

= 0 . 

dx 

du dx dv dx 

dy 

du dy 

dv dy 


Equations (7-8) are linear in du/dx, du/dy , dv/dx , and dv/dy . If 


J(u,v) m 


dF 

dF 

du 

dv 

dG 

dG 

du 

dv 


the partial derivatives in question can be determined from (7-8) by 
Cramer’s rule. 

A special case of Eqs. (7-6) is useful in applications. Let 

x = 

y = g(u,i >). 

Differentiating these equations with respect to x and remembering that 
x and y are independent variables, one obtains 

df du df dv 

1 «— — + 

du dx dv dx 

dg du dg dv 

du dx dv dx 


(7-9) 



240 FUNCTIONS OF SEVERAL VARIABLES 

These equations can be solved for du/dx and dv/dx if 


J{u,v) a 


[chap. 3 


Example 1. Let 


df 


du 

dv 


dg 

du 

dv 


u 2 — t> 2 -f- 2x « 0, 
uv — y — 0. 


Differentiating with respect to x, 

_ / du dv \ 

du &V 

V b u — * 0. 

dx dx 

du u dv v 

Hence — * 5 5 , - «* — 

dx u 2 -f v 2 dr u 2 -f v 2 

Differentiating the first of these results with respect to x gives 

du , 0 . 0t / du dv\ 


u 2 + v 2 ) + 2 -f v u 


dx 2 (u 2 4 - V 2 ) z 

u(u 2 4- v 2 ) — 2 u(u 2 — tr) uOv 3 — u 2 ) 

~ ’tfT+fiC " tf~Tv 2 ?' 

One obtains similarly dV/dx 2 , d 2 u/dx dy y and higher derivatives. 
Example 2 . Let 

x « u 4- 
(a) 

y ** 3u 4* 2e. 

Differentiating with respect to x, 

du dv 
1 - 7~ 4* 


^ du dv 

0 « 3 b 2 — » 

dx dx 


so that 

It is easily checked that 


dv 

-- * 3. 

dx 


Equations (a) can be solved for u and v in terms of x and y, and the result is 


u *» — 2 x 4 “ y, 
v * 3x ~ y. 



SEC* 7 ) THE TECHNIQUE OF DIFFERENTIATION 241 


Regarding u and v as the independent variables and differentiating these equations with 
respect to u t one finds 


1 


— 2 — + — > 
du du 


0*3 


dx 

du 


du 


Hence, 


ax 

du 


I, 


dy 

du 


« 3. 


This agrees with the result obtained by direct differentiation of (a), as of course it should. 
Note that du/dx and dx/du are not reciprocals. 

Example 3. If w * uv and 

u 2 4* e 4~ x * 0, 

(« 

v l u — y ** 0, 

one can obtain dw/dx as follows: Differentiation of ir with respect to x gives 

tKr du 

— » u - + i» — 

dx dx dx 


The value* of du/dx and dv/dx can be calculated from (b) as was done in Example I. 
The reader will check that 

dw u 4~ 2v l dw 2u l — v 

dx 1-f" 4we dy 1 4- 4m> 


PROBLEMS 

1. If u 2 4- v 2 4" V 2 — 2x * 0, u 3 4- e 3 — x 3 4- 3 y * 0, find du/dx, dv/dx , du/dy , and 
dv/dy. 

2. Find dw/dx and dw/dy if w * u/v , 


and 


j*u + v , 

y * 3u 4“ 2i;. 


3. Show that if f(x,y,z) * 0, then {dz/dx)(dx/dz) * 1 and {dx/dy){dy/dz)(dz/dx) * — 1. 
Note that in general dz/dx and dx/dz are not reciprocals. 

4. If x * x(u,v) t y * y(u,t/) with dx/du * dy/dv, and dx/dv * — dy/du } then 

+ = (*z + [ /^V+ C -Vl • 

du* dv 2 \dx 2 dyv L Vdu/ \dr/ J 


3. Show that the expressions 


Pi - 



and 


l r 2 


d 2 Z d 2 Z 

Ox* + dy*’ 


upon change of variable by means of x 


r cos 8, y * r sin B, become 



242 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 


and 

8. Show that 


Vi 


V 2 



d 2 Z 1 d 2 z 1 dz 

dr* + 7* d6* + r ir 


d*V_ 2 
* f dx* 


if P /(j? -h Cl) -f ff(x 
second derivatives. 

7. Show that 


ci), where / and g arc any functions possessing continuous 


tl , *?. = /*»_ , ^I\ 

d.r 2 dy 2 \ dr 2 dO 2 ) 


if x « e r cos 0, y * c r sin 0. 
8. Find du/ dx if 


9. Prove that 


U 2 _ t , 2 — j : 3 -f 3 y » o, 
w ~f e> — ?/ 2 — 2.r - 0. 


du dy dr; dy 
dx du dx dv 


if Fix^y^v) * 0 and G{i,y,u,v) = 0. 

10. If Vi(x,y,z) and V 2 {x,y,z) satisfy the equation 


V 2 T 


d 2 V d 2 V d 2 K 
dx 2 + dy 2 + l)z 2 


0, 


then 


u - Fi(i,y,*) + (J- 2 + + **)K,(jr,jr,*) 


satisfies the equation 


where 


v 2 v 2 r « o, 


V 2 


d 2 a 2 a 2 

ttr 2 ^ d.y 2 dz 2 


11. To indicate explicitly the variables entering in the Jacobian 


J(u,v) 


dx dy 
du du 
dx dy 
dv dv 


one frequently writes J(u,v) 



The Jacobian 




du 

dv 

dx 

dx 

du 

dv 

dy 

dy 


of the transformation (7-3) is written as J(x,y) ** J (—] < 

\%,y/ 


Prove that: 



APPLICATIONS OP DIFFERENTIATION 


243 


SEC. 8] 




where u - «(x,y), v « *(x, 2 /), x - x(£,t>), and y « y(£, v ). Hint: Write out the Jaeobians 

and multiply. 

APPLICATIONS OF DIFFERENTIATION 

8. Directional Derivatives. Formula (4-2) has a simple geometrical 
meaning when interpreted as the space rate of change of a given function 
u(x,y). Thus, let u(x,y) be specified along a smooth curve C with para- 
metric equations _ 

j — x(s), 

( 8 - 1 ) 

v - v(*), 

where s is the arc-parameter measured along C. By virtue of Eqs. (8-1), 
u(x,y) can be regarded as a function of s and the rate of change of u(x,y) 
along C is 

dn du dx du du 

— - — - + - v* (8-2) 

ds dx ds dy ds 

At a given point Po(x 0 ,yo) on C, Eq. (8-2) yields 


Ux(xoSJo) cos a + u y (x 0 ,yo) sin a, 


since dx/ds = cos a and dy/ds * sin a, as is clear from Fig. 5. It follows 
from (8-3) that the rate of change of 
u(x,y) at a given point depends only on 
the direction of the curve passing through 
that point. If the direction of C is that 
of the x axis, the angle a — 0 and du/ds 
= du/dx ; if the direction of C is that of 
the y axis, a — w/2 and du/ds *= du/dy. 

For an arbitrary direction specified by a, 

Eq. (8-3) defines the directional derivative 
of u(x,y) in that direction. Thus the de- 
rivatives u x and u v are directional deriva- 
tives in the directions of the coordinate Fig. 5 

axes indicated by the subscripts. 

We now ask the question: What is the angle a for which the directional 
derivative of u{x f y) at a given point has a maximum value? Since a 
necessary condition for a maximum is the vanishing of the derivative of 
(8-3) with respect to a, we get the equation 




244 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 


—Ux(x 0 ,!/o) sin a + u y (x 0 ,yo) cos a = 0, 
from which wo conclude that when u x (x 0 ,yo) 9* 0, 

A Uv(x 0 ,y 0 ) 

tan a = 

u x (x o,Vo) 


( 8 - 4 ) 


Accordingly, there are two values of a differing by 180° which satisfy the 
condition (8-4). The corresponding values of cos a and sin a in (8-3), 
therefore, are 


cos a 


dr VL X 



sin a = 


dr U y 

'Vul + U 2 u 


( 8 - 5 ) 


The substitution in (8-3) of the values from (8-5) with the plus sign 
the desired maximum 



= V'u* -f ub 


yields 

( 8 - 6 ) 


while the other pair of values in (8-5) gives a minimum 



- Vul + uj- 


The vector pointing in the direction of the greatest rate of 
u(x,y) at a given point (x,y) and whose length is determined 


called the gradient , and 



is called the normal derivative . 1 


increase of 
by (8-6) is 

We denote 


the normal derivative by dafdn and write 


da r-~ ~ 

— - Vu 2 x + vi- (8-7) 

dn 


A similar discussion can be applied to a differentiable function u{x,y y z) 
defined along a space curve C with parametric equations 

x ~ x(n), 


V * 0(s), 

Z «* z(s). 

We get 

du du dx du dy du dz 

ds dx ds dy ds dz ds 

1 The reason for this terminology is given in Chap. 5, Sec. 3. 


( 8 - 8 ) 


( 8 - 9 ) 



SEC. 8] APPLICATIONS OF DIFFERENTIATION 245 

where (see Fig. 6) 

dx dy dz 

— * cos (*,«), — « cos (y,s), — = cos (z,s) (8-10) 

ds ds ds 

are the direction cosines of the tangent line T at a given point P of C. 

To determine the particular direction yielding a maximum of (8-9) at a 
given point P(xo,y 0 ,Zo), we must maximize the resulting function of the 



Fig. 0 


direction cosines in (8-9). This problem, involving the determination of 
a maximum of functions of several variables, is discussed m Examph 3 ; 


Sec. 10, where it is shown that 


for (8-9) is given by the formula 


U 2 y + Vil , 


analogous to (8-7). The expression (8-11) is called the normal derivative 
of u. 


Example 1. Find the directional derivative for u{x,y) * x 2 -f y 2 at (1,1) in the direc- 
tion making the angle of 30° with the positive x axis. 

Formula (3-3) yields 

— I -2*1 cos 30° + 2y\ sin 30° - V3 + 1. 
ds l(u) la.i) i(U) 


The normal derivative at this point, as found with the aid of formula (8-7), is 

^ - V (2*) 1 + (2k)* I - 2 V2 
an l(i,D 


and the corresponding angle a, as follows from (8-4), is 45°. 

Example 2. Find the directional derivative of u(x f y,z) * xyz at (1,2,3) in the direc- 
tion of the line making equal angles with the coordinate axes. Since the angles are 
equal and the sum of the squares of the direction cosines is 1, we conclude from 

cos 2 Cm) *f cos 2 (y,s) -F cos 2 (z,s) « 1 



246 FUNCTIONS OF SEVERAL VARIABLES [CHAP. 3 

that cos ( x t s ) « cos ( y,s ) *» cos (z,$) * l/\/3. Also, at the point (1,2,3) we have 


du . du 

yz ** 6, -~ 

dx dy 

The substitution in (8-9) then yields 

du 

ds 


*■ zz ** 3, 


du 

~dz 


» xy — 2. 


6 3 2 11_ 

V3 + V3 + Vs ~ Vs 


Example 3. Show that the directional derivative of u(x,y ) in two noncollinear direc- 
tions determines the derivative in all directions. 

Let the derivative be given for directions «o and aj, so that 

u x cos «o ~h Uy sin ao =» a, 

u x cos a\ -f Uy sin a j * b, 

where a and 6 are known. If these are regarded as equations for the unknowns u z and 
the coefficient determinant is 


cos ao sin ao 1 

| j *= COS ao Sin a\ — COS aj Sin ao. 

I cos ai sin aj | 

This reduces to sin (ai — ao), which is zero only if the two directions are collinear. Hence 
u x and tty can be found, and the directional derivative is determined for every direction 
by (8-3). 


PROBLEMS 

1. Find the directional derivative of f(x,y) ** x?y 4 sin xy at (l,»/2), in the direction 
of the line making an angle of 45° with the positive x axis. 

2. Find 

I-VSHI ) 5 

if x » r cos 0, y ■* r sin 0, and / is a function of the variables r and 0. 

3. Find the directional derivative of f{x y y ) * x*y 4 e vx in the direction of the curve 
which, at the point (1,1), makes an angle of 30° with the x axis. 

4 . Find the normal derivative of u » j : 2 4 y 1 4 z 2 at the point (1,2,3) and the direc- 
tional derivatives at that point along the line joining (0,0,0) and (1,2,3). 

9. Maxima and Minima of Functions of Several Variables. A function 

f{x y y) defined in a region R is said to have a relative maximum at a point 

(o,6) if A/ = f(a + h,b + k)~ f(a,h) < 0 (9-1) 

for all values of h and k in the neighborhood of ( a,b ). It is said to have a 
relative minimum at (a,b) if 

Af s f(a 4- h,b 4 k) — f(a, b) > 0 (9-2) 

for all values {h,k) in the neighborhood of (a, b). 

The requirement that the inequalities (9-1) and (9-2) hold for all values 
(h,k) in the neighborhood of (a,b) implies that we are concerned here 



APPLICATIONS OF DIFFERENTIATION 


247 


SEC. 9] 

only with the interior and not the boundary points of the region. A func- 
tion may attain a maximum or a minimum value on the boundary of the 
region, but the behavior of functions on the boundary requires a separate 
investigation, the nature of which will be clear from the sequel. The 
greatest and least values assumed by f(x 9 y) in the closed region are called, 
respectively, the absolute maximum and the absolute minimum. In the 
following discussion we dispense with the adjective “relative,” and we 
shall refer to relative maxima and minima simply as maxima and minima. 

Let it be assumed that f(x,y) attains a maximum (or minimum) at some 
interior point (a, b). Then the func- 
tion f(x,b) of the variable x must 
attain a maximum (or minimum) at 
x ~ a. From the study of functions 
of one variable it follows that the 
derivative of f(x,b), if it exists, must 
vanish at x = a. The derivative may 
cease to exist at the critical points 
when the behavior of the function is 
like that shown in Fig 7 in the neigh- 
borhood of x — ait x “ a 2i and x = c/ 3 . 
a maximum (or minimum) of f{x,b) at x 



Thus, a necessary condition for 
= a is that 


d[ 

dx 


0 


(9-3) 


if this derivative exists at x — a. 

A similar consideration of the function /(a,y) leads to the conclusion 
that 

df 

— = 0 at y = b ( 9 - 4 ) 

dy 

whenever this derivative exists. 

The coordinates (a, b) thus satisfy the pair of equations 



df 

dy 


(9-5) 


at any point ( a,b ) where /(a*, y) attains a maximum or minimum. 

This discussion is capable of extension to functions of any number of 
variables to yield a theorem. 

Theorem. A function f(x i,x 2 , . . . f x n ) of n independent variables x t attains 
a maximum or a minimum only for those values of the variables Xifor which 
fx v fx it . . fz n either vanish simultaneously or cease to exist. 

We emphasize that the conditions stated in this theorem are necessary 



248 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 

but not sufficient for a maximum or a minimum. 1 Although the matter of 
sufficiency can usually be determined from the nature of the problem and 
from physical considerations leading to its formulation, we record here a 
test that may prove useful to settle doubtful cases. 2 

If /Or,?/) is a function with continuous second partial derivatives, and if 
/*(<!,&) ~ 0 and f y (a,b) ~ 0 , then /(a, b) is a maximum provided that 

D = fl y (a,b) - fxx(a,b)fy V {a,b) < 0 

and f TX (o i t b) < 0, f vy (a } b) < 0; it is a minimum if D < 0 and f xx {a,b ) > 0, 
>0; it is neither maximum nor minimum (a saddle point) if 
1) > 0. This test gives no information if 1) = 0, just as the condition 
/"(a) — 0 gives no information for the function }{x) with /'(a) — 0. 

Before proceeding to a further study of maxima and minima, we give 
two examples illustrating the developments of this section. 


Example 1. A long piece of tin 12 in 



wide ib made into a trough by bending up 
the sides to form equal angles with the base 
(Fig. 8). Find the amount to be bent up 
and the angle of inclination of the sides that 
will make the carrying capacity a maximum. 

The volume will be a maximum if the 
area of the trapezoidal cross section is a 
maximum The area is 

A *» I2x sm 0 — 2x 2 sin sm 6 cos 6, 


for 12 — 2x is the lower base, 12 — 2-c -F 2x cos 6 is the upper base, and x sin 6 is the 
altitude. Then, 

dA ,, ,, „ 

— « 12a- cos 6 — 2j" cos Hr cos* ft — x l sirr 6 
06 

— r( 12 cos 6 — 2x cos 6 ■+* x cos ; 0 — x sin 2 6) 

, 0A . , 

and — — 2 sm 0 ( 0 — 2x + x cos 6). 

Ox 


Now OA/Ox * 0 and dA 1 06 « 0 if sin 0 — 0 and x «* 0, which, from physical considera- 
tions, cannot give a maximum. 

There remain to be satisfied 

6 — 2.r -f x cos 6 «= 0 


and 


12 cos 6 — 2x cos 6 + x cos 2 6 ~ x sin 2 6 ~ 0 


Solving the first equation for x and substituting in the second yield, upon simplification, 

cos 6 * or 6 — 00°, and x * 4. 

Since physical considerations show that a maximum exists, x ~ \ and 0 ~ 60° must 
give the maximum. 


1 Recall, for example, the situation when f(x) has a point of inflection with a horizontal 
tangent, 

, * A proof and further discussion are contained in I. S. Sokolnikoff, “Advanced Calcu* 

' hi»/' sec, 89, McGraw-Hill Book Company, Inc., New York, 1939. 



249 


SEC. 10 ] APPLICATIONS OF MFFEHENTIATXOM 

Example 2. Find the maxima and minima of the surface 

x 2 i i L 


- ?5 - 2 <*. 


Now, 

which vanish when x « y 

* 

dr 2 * 


6 2 


dz 1 2 

dx c a 2 

* 0. But 

_I_ dh 

ah' dy 2 


to 

dy 


Wc 


1 V 
cb *’ 


d»g 
dx dy 


0. 


Hence, D ==> l/a 2 & 2 c 2 , and consequently, there is no maximum or minimum at x =* y « 0. 
The surface under consideration is a saddle-shaped surface called a hyperbolic paraboloid. 
The points for which the first partial derivatives vanish and /) > 0 are called minimax . 
The reason for this odd name appears from a consideration of the shape of the hyperbolic 
paraboloid neur the origin of the coordinate system. The reader will benefit from sketch- 
ing it in the vicinity of (0,0,0). 


PROBLEMS 


1. Divide a into three parts such that their product is a maximum. Test by using 
the second-derivative criterion. 

2. Find the volume of the largest rectangular parallelepiped that can be inscribed in 
the ellipsoid 


a* ^ b* ^ c* 


= 1. 


3. Fmd the dimensions of the largest rectangular parallelepiped that has three faces 
in the coordinate planes arid one vortex m the plane 


x y z 

~ + 7 -f “ 

a b c 


1. 


4. A pentagonal frame is composed of a rectangle surmounted by an isosceles triangle. 
What are the dimensions for maximum area of the pentagon if the perimeter is given as P? 

5. A floating anchorage is designed with a body in the form of a right-circular cylinder 
with equal ends that are right-circular cones. If the volume is given, find the dimensions 
giving the minimum surface area. 

6. Given n points P, whose coordinates are (x if yi,z t ) (i ~ 1,2, . . ., n). Show that the 
coordinates of the point P(x,y } z ), such that the sum of the squares of the distances from 
P to the P* is a minimum, are given by 



10. Constrained Maxima and Minima. The discussion in the preceding 
section was confined to the calculation of the maximum and mi nimum 
values of functions of several independent variables. In a large number 
of investigations, it is required that the maximum and minimum values 
of a differentiable function f(z . . ,x n ) be found when the variables Zi 
are connected by some functional relationships, so that the %i are no longer 



250 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 

independent. Such problems are called problems in constrained maxima 
to distinguish them from the problems in free maxima discussed in Sec. 9. 

To avoid circumlocution, we shall speak of the maximum or minimum 
values as the extreme values. Thus, let us consider the problem of finding 
the extreme values of the function 

u « f(x,y,z ), (10-1) 

in which the variables x, y f z are constrained by the relation 

*p(x,y>z) ~ 0 . ( 10 - 2 ) 

This problem can be solved by the procedure of Sec. 9 as follows: Suppose 
that the constraining relation (10-2) is solved for one of the variables, 
say z f to yield a differentiable function 

s « (10-3) 

If one substitutes z from (10-3) in (10-1), there results the function 

y = f[x,yMx,v)] - F(*,y) ( 10 - 4 ) 

of two independent variables x , y to which the considerations of Sec. 9 
apply. 

However, either the solution (10-3) may be difficult to obtain or the 
function F(x,y) in (10-4) may be so unwieldy that the simultaneous equa- 
tions F x (x,y) — 0, F y (x,y) — 0 are unpleasant to deal with. In this 
event an ingenious method devised by the great F reneh analyst Lagrange 
often leads to a manageable* and symmetric system of equations for the 
determination of extreme values. The central idea of the method hinges 
on the following observation. In Sec. 9, we saw that a necessary condition 
for a relative extremum of the differentiable function fixity. . . ,x„) of 
n independent variables is the simultaneous vanishing of all partial deriva- 
tives f Xt . Inasmuch as the total differential of /is 

df = f Xl dx i + } Xt dx 2 H h fx n dx n , 

it is clear that df = 0 whenever each f x% ■= 0. Conversely, if df = 0, the 
partial derivatives f x% vanish, since the dx % are independent. But it is 
also true that the vanishing of the total differential is a necessary condi- 
tion for an extremum of f(x i,T 2 ,...,x n ) even when the variables x t are 
dependent because of the invariant character of df stated in the theorem 
of Sec. 5. We can thus state a theorem: 

Theorem. A necessary condition for an extremum of a differentiable func- 
tion f(x h X 2 , . . . ,x n ) is the vanishing of its total differential at the maximum and 
minimum points of the function. 

We proceed now to a discussion of the method of Lagrange multipliers 
for determining the extreme values of the function in (10-1) subject to 
the equation of constraint (10-2). 



SBC. 10] APPLICATIONS OF DIFFERENTIATION 251 

By the theorem just stated, the differential of (10-1) vanishes at the 
critical points so that 

df df df 

— dx + —dy + -dz~ 0. (10-5) 

dx dy dz 

Also, since <p(x } y,z) — 0, its total differential vanishes and we can write 


dtp dtp dip 

~ dx H dy H dz = 0. 

dx dy dz 


( 10 - 6 ) 


Let Eq. (10-6) be multiplied by some parameter X and then added to 
(10-5). The result is 


(* + + (* + X *U + + X s ')* , 0. 

\dx dx/ \dy dy/ \dz dz) 


(10-7) 


if we regard x and y as independent variables, and suppose that d<p/dz ^ 0 
at the point where the extremum is attained, then we can find a X such 
that at this point 


df dip 

— + X — 
dz dz 


= 0 . 


(10-8) 


With this choice of X, Eq. (10-7) reduces to 


/ df d<p\ / df dip\ 

( b X — ) dx l f* X — ) dy = 0. 

\dx dx) \dy dy) 


But since dx and dy are independent increments, we conclude from this 
equation that 

df d<p 

~ + X— = 0, 
dx dx 


dy 


dip 

+ X — - 0. 
dy 


(10-9) 


The system of three equations (10-8) and (10-9) contains four unknowns 
x, ?/, z , X, and we must adjoin to it the fourth equation (10-2) to obtain the 
complete system for the determination of the unknowns 

If dip/dz = 0 at the point where the extremum is attained, but dip/dy y* 0, 
the roles of z and y in the foregoing discussion are interchanged. Clearly, 
the method will fail to yield the desired value of X when <p Xf <p yy and <p * 
vanish simultaneously at the point where f{x,y,z) has an extremum. 

Before proceeding to extend the Lagrange method to the study of 
extreme values of functions with several constraining conditions, we con- 
sider four instructive examples. 



252 FUNCTIONS OF SEVERAL VARIABLES [CHAP. 3 

Example 1. Find the maximum and the minimum distances from the origin to the 
curve 

5x 2 4 fay 4 by 1 - 8 * 0. 

The problem here is to determine the extreme values of 

f(x,y) « x 2 4 y 2 

subject to the condition 

<p(x } y) ss 5x 2 -f 6xjy 4 5j/ 2 - 8 « D. 

Equations (10-9) and (10-2) in this case read 

2x 4 X(U)x -f f>y) « 0, 

2 y 4 X(6x 4 10 y) - 0, 

5x 2 4* fay 4 by 2 — 8 * 0. 

Multiplying the first of these equations by y and the second by x and then subtracting 
give 

6 X(y 2 — x 2 ) «• 0, 

so that y ® rkx. Substituting these values of y in the third equation gives two equations 
for the determination of x, namely, 

2x 2 « 1 and x 2 = 2. 

The first of these gives / &e x 2 4- ?/ 2 — 1, and the second gives / e- x? 4* ?/ 2 * 4. Obvi- 
ously, the first value is a minimum, whereas the second is a maximum. The curve is an 
ellipse of semiaxes 2 and 1 whose major axis makes an angle of 45° with the x axis. 

Example 2. Find the dimensions of the rectangular box, without a top, of maximum 
capacity whose surface is 108 in. 2 
The function to be maximized is 

f(x t y,z) rn xyz , 

subject to the condition 

xy 4 2xz 4 2 yz « 108. (10-10) 

Equations (10-8) and (10-9) yield 

yz 4- My 4 2 z) ® o, 

xz 4 X(x 4 2 z) * 0, (10-11) 

xy 4- X(2x 4 2 y) « 0. 

In order to solve these equations, multiply the first by x, the second by y } and the last 
by 2 , and add. There results 

X(2 xy 4 4xz 4 4yz) 4 3 xyz « 0, 

or X (xy 4 2xz 4 2 yz) 4 %xyz » 0. 

Substituting from (10-10) gives 

108X 4 Hxyz * 0, 


or 


X 


xyz 

72 



BMC. 10 ] APPLICATIONS OF DIFFERENTIATION 

Substituting this value of X in (10-11) and dividing out common factors give 


253 


1 -~(y+2z) -o, 

1 -~(*+W -0, 

1 _ -L (2z + 2 y) -0. 

From the first two of these equations, it is evident that x » y. The substitution of x « y 
in the third equation gives z * 18 /y. Substituting for y and z in the first equation yields 
x » f>. Thus, x ** 0, y =» 6, and 2 * 3 give the desired dimensions. 

Example 3. Show that the maximum value of the directional derivative of u(x,y t z) 
at any point is given by 

~ - V^TufT4. 

We write the directional derivative [see Eq. (8-9)] in the form 

du 

f(<x,P,y) 22 ~ « u x cos a ~f ii v cos /3 4- u g cos y, (10-12) 


where cos a « cos (r,s), cos# * cos(y,«), cos 7 - cos (2,8), and maximize /(a, 0,7) sub- 
ject to the constraining condition 

*>(«,£, 7) 23 cos 2 a -f- cos 2 /9 4~ cos 2 7 — 1 =» 0. (10-13) 

The system of Eqs (10-8) and (10-9) then yields 


— u x sin a - 2\ cos a sin a «* 0, 

— Uy sin /3 — 2X cos 0 sin P = 0, (10-14) 

— u* sin 7 — 2X cos 7 sin 7=0. 

The case when either sin a, sin j8, or sin 7 vanishes is trivial because of the constraining 
condition (10-13). Thus, the system (10-14) reduces to 

u x =» 2X cos a, Uy * 2X cos u t ** 2X cos 7, (10-15) 

and we conclude that 


4 + 4 + u\ 


4X 2 . 


Thus, X * 4- wj + uf, and the substitution of this value of X in (10-15) gives 


cos 0 


Vu| 4~ a 2 4- w? + *4 + u l 

On inserting these values in (10-12) we get the desired result 


cos 7 


u t 


Vu'i 4- U 2 V -F u\ 



254 


FUNCTIONS OF SEVERAL VARIABLES 


Example 4. Find the shortest distance from the origin to the curve y 
in Fig. 9. We apply the procedure employed in Example 1 to minimize 

f{x,y) - x 2 4- y 2 

subject to the constraining condition 

<p(*,y) m if - (x - l) 8 « 0. 
Equations (10-9) now yield 

2x - 3X(x - l) 2 - 0, 

2 y -f 2 Ay * 0, 


[chap. 3 

« (x ~ D* 
(10-16) 

(10-17) 

(10-18) 


which must be solved together with (10-17). The 
system (10-17) and (10-18) has no solutions for#, y , 
and X This becomes obvious on noting that the mini- 
mum is attained at x » 1, y * 0, and if we insert 
these values in (10-18), the first of the resulting equa- 
tions yields a nonsensical result 2=0 while the second 
is true for all values of X. The reason that the La- 
grange method this time has failed to give the solution 
is simple. The method depends ou the assumption 
that not both <p x and <p v vanish at the point where 
the extremum is attained. In our case *>*(1,0) = 0 
and <p„(1,0) = 0. The moral of this example is that the Lagrange method yields the 
solution of the problem only when the system of Eqs. (10-8) and (10-9) can be solved 
for X. 



PROBLEMS 

1. Work Probs. 1, 2, and 3, Sec. 9, by using Lagrangian multipliers. 

2 . Prove that the point of intersection of the medians of a triangle possesses the prop- 
erty that the sum of the squares of its distances from the vertices is a minimum. 

3 . Find the maximum and the minimum of the sum of the angles made by a line from 
the origin with (a) the coordinate axes of a cartesian system, (b) the coordinate planes 

4 . Find the maximum distance from the origin to the folium of Descartes x s -f y 3 - 
3 axy = 0 , 

6. Find the shortest distance from the origin to the plane 

ax -p by -f cz ** d. 

11. Lagrange Multipliers. We now extend the considerations of Sec. 10 
to cases where the extremum of the function f(zi,x 2 , . . . ,x n ) is sought under 
several conditions of constraint. 

We consider first the function 

w = f(x,y,u,v), (11-1) 

in which the variables are constrained by two relations 

<Pi{x,y,u,v) = 0, 

V 2 (x,y,u,v) = 0. 


(11-2) 



SEC. 11] 


APPLICATIONS OF DIFFERENTIATION 


255 


If w takes on the extreme values for certain values of (x,y } u,v), then for 
such values 

df df df df 

— dx -j dy d du dv = 0, (11-3) 

dx dy dudv 

by the theorem in the preceding section. Also, (1 1-2) yields two equations; 

dtp 1 d<pi dtp i d<p\ 

~~dz + — -dy + ~^du + — dv - 0 
dx dy du dv 


d<p2 dtp 2 dtp2 dtp2 

dx - — dy "|~ — '■*' du -f* — dv — 0. 

dx dy du dv 


(11-4) 


We multiply the first of these by Xi and the second by X 2 , add the results 
to (11-3), and obtain 


Now, if 


d<P2\ 

J , (df 


d<pi 


+ x 2 — ) 

dx + ( — 

+ x 

-L \ 2 J dy 

dx) 

Va y 


a?/ 

dy/ 


(df 

dtp i 

d(p2\ 

+ 

(— + x, 


+ x 2 - 

— )dv 


\du 

du 


du) 



/df 

dtp\ d<pz\ 


+ 

l- + x,- 

— + Xo — dv 



\dv 

dv dv / 



dtpi 

d<Pi 




du 

dv 



J(u,v) = 



* 0, 



dtp 2 

CS 1 

<T> | 




du 

dv i 



0. (11-5) 


the values of Xj and X 2 can be found such that 

df dtpi dtp2 

i + Xi +X =0, 

du du dxi 

df dtpi dtp 2 

~ + h + x 2 — - 0, 

dv dv dv 


( 11 - 6 ) 


and accordingly (11-5) reduces to the sum of two terms involving arbitrary 
differentials dx and dy. The fact that they are arbitrary enables us to 
conclude that 

df dtpi dtp 2 



256 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 

The system of six equations (11-6), (11-7), and (11-2) serves to determine 
the parameters Xi, X 2 and the point (x,y;u,v) at which the extreme is at- 
tained. 

The foregoing procedure may be extended to cover the case of more 
than two constraining conditions and we obtain the following rule: 

Rule. In order to determine the extreme values of a function 


f(xi,x 2t . . .,x„) 

whose variables are subjected to m constraining relations 

(11-8) 

^i;2r*>;») =0, i * 1, 2, . . m, 

form the function 

m 

F = / + X,v>» 

»»i 

(11-9) 

and determine the parameters X, and the values of Xi, x 2 , . . . 
equations 

dF 

, x n from the n 

— = 0, j - 1, 2, 

dXj 

(11-10) 


and the m equations (11-9). 

It should be carefully noted that the applicability of this rule to specific 
problems depends on the possibility of determining the multipliers X». 
The existence of the X x was established above only under the hypothesis 
that J 0. 


Example: As an illustration, consider the problem of determining the maximum and 
the minimum distances from the origin to the curve of intersection of the elUjasoid 

x 2 v 2 z 2 
l i - j 3 } 

a 2 ^ 6 2 * c* 

with the plane 

Ax *f By -f Cz * 0. 

The square of the distance from the origin to any point ( x,y,z ) is 

/ * x 2 + y 1 + * 2 , 

and it is necessary to find the extreme values of this function when the point (x,y t z) is 
common to the ellipsoid and the plane. The constraining relations are, therefore, 


x 2 y 2 z 2 
a 2 ^ b 2 ^ c 2 


1 


(a) 
and 

(b) ip 2 m Ax 4" By ■+* Cz » 0. 

The function F ** f - f -f Mw is, in this case, 

*£ z? 


F - x 2 + y 2 + z 2 + Xi ^ ~ - l) + 2 MAx + By + Cz), 



APPLICATIONS OP DIFFERENTIATION 


SBC. 12] 


257 


where the factor of 2 is introduced in the last term for convenience. Equations (11-10) 
then become 

x 4“ Xi 4- XsA “ 0, 

or 


(0 


V + Xi ^ + XiB — 0, 


* -f* Xi -r -f- X 2 C *» 0, 

cr 


These equations, together with (o) and ( b ), give five equations for the determination 
of the five unknowns x, y, z, Xi, and X 2 . If the first, second, and third of equations (c) 
are multiplied by x , y, and z, respectively, and then added, there results 

o* + h + ?) + ^ (Ax + By + Cz) ■ °- 

Making use of (a) and (6), it is evident that 


X, = -(z* + y* + z *) - -/. 
Setting this value of Xj m ( c ) and solving for x } t/, and z, 


y + XjB = 0, 

* 0 - i) + x * c - °- 


X 2 A a 2 


n? — 


f 


\ z Bb 2 

MCc 2 


c 2 — / 

When these values of x t y f and z are substituted in (b), one obtains 

yy By cy 

a 2 — / 6 2 — / + r 2 ~ / 

from which /can be readily determined by solving the quadratic equation inf. 


PROBLEMS 

1. Find the point P, in the plane of the triangle ABC , for which the sum of the dis- 
tances from the vertices is a minimum. 1 

2. Find the triangle of minimum perimeter which can be inscribed in a given triangle. 

12. Taylor’s Formula for Functions of Several Variables. Let f(x 9 y) be 
a function of two variables x and y that is continuous in the neighborhood 
of the point (a, 6) and that has continuous partial derivatives, up to and 
including those of order n, in the vicinity of this point. 

1 See E. Goursat’s “Mathematical Analysis/' English ed., vol. 1, p. 130, for a detailed 
discussion of this interesting problem. 



258 FUNCTIONS OF SEVERAL VARIABLES fcHAP. 3 

If a new independent variable t is introduced with the aid of the relations 

x = a + at, y = b + fit, (12-1) 

where a and fi are constants, a function of the single variable t will result, 
namely, 

m - f(x,y) = f(a + at, b + 0t). (12-2) 

Expanding F(t) with the aid of the Maelaurin formula gives 

F"( 0) , F in) (dt) 

F(t) = F{ 0) + F'(0)l + — ^ t 2 + • • • + t n , (12-3) 

2! n! 

where 0 < 8 < 1. 

It follows from (12-1) and (12-2) that 1 

dx dy 

F'(t) =f*(x,y)-+f v (z,y)- 

dt dt 

* fx(x,y)a + f y (r,y)p. 

Calculating F"(t) and from this expression gives 

dx dy 

F”(t) - [fxz(r,y)a + fyz(x,y)p] — + lfxp(x,y)a + jf w 0r,y)/3] ~~ 

dt dt 


, = fzx(x,y)a 2 + 2 f zy (x,y)a0 + f m (x,y)0 2 , 

and 

„ „ dx 

F"'(t) - Uxxx{x,y)a 2 + 2f xvz (*,y)<xf} + f n *(T,y)P 2 ]- 


lfxxy(x,y) ci “H 2f zyy (x,y)a0 -f- f yyy {x,y)0] 

at 

= fxrx(x,y)a 3 + 3f zzy (x,y)a 2 0 + 3f zyl/ (x,y)a0 2 + f yyy (x,y)0*. 

Higher-order derivatives of F[t) can be obtained by continuing this 
process, but the form is evident from those already obtained. Symbolically 
expressed, 


F'(t) =(a~ + P Pl f(x,y) 
\ dx dy/ 


1 d 

d N 

2 

k 

F'\t) -(«- 

+ 18 — 


\ dx 

dy/ 

f 

{ d 

d\ 

3 

i 

F”'(t) = (a- 

+ 0 — 

J f(x,V) 

V dx 

dy) 



df df 

dx dy 


— ,_2 


d 2 f 


d 2 f 


or + 2 a|3 - — ~ + 0 * 


dx 1 

d*f 


a" — - + 3 c?0 


dx dy 

d*f 


dx* 


dx 2 dy 
+ 3ajS 2 


dy 2 

d*f 


dx dy 2 


+ , 


d*f 

dy* 


1 See Sec. 4. 



SBC. 12] 
Then 
F lB, (<) = 


APPLICATIONS OP DIFFEKBNTIATION 


259 


a a\ n a n f a n f 

T + P T ) f(x,y) m a" — + ^ 

dx by! dx n dx n 1 < 


dx n dx n ^dy 

d n f d n f 

+ • • • ■ + —4^ + r -L 

dx by 1 dy n 


where 


r!(n — r)! 

Since t — 0 gives x — a and y = b, 

F( 0) - f(a,b), F’{ 0) = af x (a,b) + 0f v (a,b), .... 

Substituting these expressions in (12-3) gives 

F(t) m f(x,y) - f(a, b) + [of Mb) + 0f y (a,b)]t 

+ l<x 2 f ZI (a,b) + 2a0f zy (a,b) + 0% y (a,b)) L + • ■ • + R n , 

Z ! 


r ( a ay 

where R„ *= - la \- 0 — ) j 

n\ \ dx du / 


f(a + Bat, b + 60t). 


Since at — x — a and 0 t — y — b, the expansion becomes 
J{x,y) = /(«,&) + f z {a,b){x - a) + f v (a,b)(y - b) 

+ ~j Uxx(a,b)(x. - a) 2 + 2 f xy (a,b)(x - a)(j/ - 6) + f vv (a,b)(y - b) 2 ] 

+ ---+R n . ( 12 - 4 ) 

This is Taylor's expansion for a function f(x,y) about the point (a,b). 
Another useful form of (12-4) is obtained by replacing x — a by h and y — b 
by k, so that x = a + h and y = b + k. Then, 

f(a 4- h,b + k) = f(a,b) 4- fx(a,b)h -f f v {a,b)k 


+ - ( JrMb)h 2 4- 2 U(a,b)hk+f uy (a,b)k 2 ] 

Z ! 


where R 


—u. 

n! \ 


d d \ n 

fl .-j- ^ Qfa* Jy _J_ 

dx du/ 


+ ---+Rn, ( 12 - 5 ) 


n! \ dx dy/ 

This formula is frequently written symbolically as 

/(a + h,b + k)- f(a,b) + (h~-{-k ~)f(a,b ) 

\ dx dy/ 


1 / d d\ 2 

+ 2\\ h 7x + k Ju) f{a,b)+ '" +Rn - 



260 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 

In particular, if the point (<t,&) is (0,0), the formula (12-4) reads 

/<*,¥) =“ /(0,0) + /,( 0 , 0 )* + f v (0,0)y 

+ “ f/xx(0,0)x 2 +2f xv (0,0)xy + f vv (0,0)y 2 ] 

+ * * * + Rn, (12-6) 

1 / d d\ n 

where R n = ~~ [x h y — ) f(0x f $y) f 0 < 6 < 1. 

n\ \ dx dy/ 


This development is known as the Maclaurin formula for functions of two 
variables. It is seen from (12-6) that the Maclaurin formula expresses 
the function f(x,y) in a series each term of which is a homogeneous poly- 
nomial in x and y. 

The procedure outlined above can be generalized easily to yield similar 
expansions for functions of more than two variables. 

Example: Obtain the expansion of tan"" 1 ( y/x ) about (1,1) up to the third-degree terms: 


y 

f(x,y) = tan 1 

X 

/(1,1) - tan- 1 

U**)" - x * + / 

/x(l,l) - - \ ; 


/»(!.') — 

, , N 2 xy 

f X z(x,y) - + 

i-i je* 

1! 

s 

V 2 - J 2 

fxyix,y) = + 

VU) = 0; 

fnU ’ V) “ (x r + !/ 2 ) 2 ’ 

«U) - - 


xX_ j-i(x-l)+id,-l)+Afl(r-l)*-i(v-l)*] 


PROBLEMS 

1. Obtain the expansion for xy 2 + cos xy about (1 ,tt/ 2) up to the third-degree terms. 

2. Expand f(x, y) — e ** at (1,1), obtaining three f orms 

3. Expand e x cos y at (0,0) up to the fourth-degree terms. 

4. Show that for small values of x and y 

e x sin y y *f xy (approx), 
u 2 

and e T log (1 + y) « y + xy - ~~ (approx). 

3. Expand f(x y y) « x z y j- x 2 y -f 1 about (0,1). 

6. Expand Vl — x 2 r — y 2 about (0,0) up to the third-degree terms. 



&BC. 13] INTEGRALS WITH SEVERAL VARIABLES 261 

7. Show that the development obtained in Prob. 6 agrees with the binomial expansion 
of [1 - (** + y 2 ))* 


INTEGRALS WITH SEVERAL VARIABLES 


13. Differentiation under the Integral Sign. The fundamental theorem 
of integral calculus states that whenever f(x) is a continuous function in 
the closed interval (a, 6) and F(x) is any function such that F'(x) = f(x), 
then 

[ Ul f(r)dz « F( Ui ) - F(u o) ( 13 - 1 ) 

for any two points uq and u } in the interval. If u 0 and u t are differentiable 
functions of another variable a, so that 


= Uo(<x), u i « Ui (<*), 

the right-hand member in (13-1) is a function of a and the chain rule gives 
dF (tq) du\ dui 

—j - 1 - F\ui) — - /(Kj) 
aa da da 


Since a similar result holds for F(u { )), differentiation of (13-1) yields the 
important formula 


d 

da 


ru^a) du\ 

/ ftfdx-ffa) — 

da 


- /(« o) 


duo 

da 


(13-2) 


If the variable a in (13-1) occurs under the integral sign, so that the 
integral takes the form 

*l«) - [ Ul J(T t a)dx % (13-3) 

Ju {) 

we can compute the derivative of <p(a) by calculating the limit of the 
difference quotient A«p/A« as Aa — > 0 This calculation is simple when 
the limits u< h // j are constant. Indeed in this case (13-3) gives 


r u i f u i 

A<^ - <p(a + Aa) — <p(a) = / f(x, a + Aa) (lx - / f(x,a) dx 

Ju„ Juo 

= f [/Or, a + Aa) - f(x ,a)] dx. 

J Uq 

Dividing by Aa and taking the limit as Aa — ► 0 give 

<p ( a + Aa) — <p(a) <* + Aa) — /(x f at) 

<p'(a) ?? lim = hm / dx 

a« *-♦ o Aa a« -+ o yw o Aa 

(13-4) 

provided the limit on the right exists. 



262 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 


If we knew that 

ru, f{x, a + Aa) 


lira /“* 


■ /(*,«) 


Aa J *t> Aa o 

then the right-hand member of (13-4) would give 

ru, df 


, r u i .. /fo a + Aa) - /(x) 
ax ~ / lim ax, 

JUn 


Aa 


(13-5) 


L 


uo da 


-dx 


by the definition of partial derivative. We could then conclude that 

d ru, ru, 

— I /(x,a) dx~ /«(x,a) dx, u 0 , iq const. (13-0) 
da Ju q Ju o 


Interchanging an integral with a limit operation as in (13-5) is not valid 
in general, 1 but the equality of (13-5) can, in fact, be justified when f a (x, a) 
is continuous, and hence (13-6) holds in that case. 

Equation (13-2) requires that the integrand be independent of a, while 
(13-6) assumes that the limits of integration are independent of a. 

When the limits and also the integrand depend on a, it can be shown 2 
that the correct formula is given by addition of (13-2) and (13-6); namely, 


«, (a) 


dux 

, , , s /(*»«) dx = /(« i»«) — 

da J u 0 (a) da 


/. 


du 0 ru,(a) 

/(« 0,«) — + / , faM dx (13-7) 

da y u 0 (a) 


provided that uo{a) and iq(a) are differentiable and /(x,a) and / 0 (x,a) 
are continuous. The formula (13-7), known as Leibniz' s formula, will now 
be illustrated by several examples. 


Example 1. Evaluate -- 
da 

(13-6) yields, when a^O, 


/ log (x 2 + a 2 ) dx. Inasmuch as the limit* are constants, 
J o 

f log (x 2 + a 2 ) dx - f dx . 

da Jo Jq x 2 -j- a 2 


The resulting integral is easily evaluated by the fundamental theorem of integral calcu- 
lus, since 

:( 2ten_, 3 


A. 

dx 1 


We thus obtain 


2a 

x 1 + a 2 


d_ 

da 


f log (x 2 + « 2 ) dx ® 2 tan 1 - I » 2 tan"*” 1 -• 
yo « lo a 


1 The reader can verify that 

f 1 1 2x 


r 1 1 2x f 1 1 

lim / 2 ~T — 2 dx ^ I lim 

o yo log a x a Jo « -♦ 0 log a 


2x 


log a x* -f 


dx. 


•See I. S. Sokolnikoff, “Advanced Calculus/’ pp. 121-122, McGraw-Hill Book Com- 
pany, Inc , New York- 1939. 



SEC. 13] 


INTEGRALS WITH SEVERAL VARIABLES 
,*K 


263 


x 2 dx , find <p'(x) first by evaluating the integral and then 
o 

differentiating, and also by the Leibniz rule. To avoid confusing the parameter x ap- 
pearing in the limits of the integral with the variable of integration x, we write 

V?(x) » f t 2 (it » ^7 3 I a 

Jo 10 

Hence *p'(x) «■ the other hand, the application of the rule (13-2) yields 


*>'(*) 


dx 


F 

Jo 


t 2 dl = (x 4 ) 2 


dx^ 

dx 


thus checking the result previously obtained. 

Example 3. Find d<p/da if <p(a) ~ / e~ xtla * dx Since the integiand and the limits 

J a 

in this integral are functions ot a, we use formula (13-7) Then 


dip 

da 


/ -« 2x i 

dx + t~ 4 (2) - e ~ at ( -2a) 

- a <* 

,2a 2 X* e _ xVal ^ 2e _ 4 2aC- aJ . 

-/-« 2 a 3 


The integral appearing in this expression cannot be evaluated in a closed form in terms 
of elementary i unctions, but it can be readily computed in infinite series (see Chap 2, 
Sec. 10) 

Example 4 Formula (13-(V) can sometimes be used to evaluate definite integ“j,ls, 
Thus consider 

r l x a - 1 

Y?(a) / - dx, a > 0. 

7u log x 

Differentiating under the integral sign, we get 

f l x a log x 

P (a) - I - dx - x a dx. 

Jo log x Jo 

The evaluation of the integial is easy, and we find 


(13-8) 


/(a) = — 


1 


Integrating again we get 


+ 1 lo a + 1 

<p(oc) = log (a + 1) + f. 


(13-9) 


To evaluate the constant c, we note that for a = 0, (13-8) gives <p{ 0) - 0 while (13-9) for 
a « 0 requires that v»(0) log 1 + r. Hence c « 0 We finally imve 


i 


log X 


dx =* log (a -f 1). 


PROBLEMS 




1. Find , 

da Jo 

result by direct calculation. 


r*72 

if <p{a) - / sin ax dx by using the Leibniz formula, and check your 
Jo 



264 


FUNCTIONS OF SEVERAL VARIABLES 


[chap. 3 


а. Find ~ if *(«) - f*( 1 - « cos i) s dx. 

da Jo 

5. Find ~ if ?(a) « f tan" 1 dx. 

4a Jo or 

4. Find if <p(a) «■ C tan (x — a) dx. 
da Jo 

б. Find if <?(z) *» f Vx dx. 

oa? do 

6. Show in the manner of Example 4 that 

*►(<*) « / log (1 + a cos x) dx « ir log 

Jo 


1 + 

"2 


7. Differentiate under the sign, and thus evaluate 

dx TV 

a — COS X (a 2 — 1) H 


U 


dx 


l 


(a ~ COS x) 2 
if a 2 > 1. 


if a 2 < 1. 


by using 


8 . Show that 


9. Verify that 


f log (1 — 2a cos x -f a 2 ) dx « 0 if « 2 < 1 
Jo 

— V log a 1 if a 2 > 1. 

1 r* 

y » ~ I f (a) sin k{x — a) da 
k Jo ' 


is a solution of the differential equation 

d 2 y 


dx 2 


+ k 2 y « f(x), 


where k is a constant. 


14. The Calculus of Variations. Physical lawn can often be deduced 
from concise mathematical principle's to the effect that certain integrals 
attain extreme values. Thus, the Fermat principle of optics asserts that 
the actual path traversed by the light particle is such that the integral 
representing the travel time between two points in every medium is a 
minimum. Also a considerable part of mechanics can be deduced from 
the principle of minimum potential energy, stating that the equilibrium 
configuration of a mechanical system corresponds to the minimum value 
of a certain integral related to the work done on the system by the forces 
acting on it. For example, the shape assumed by a flexible chain fixed 
between two fixed points is such that its center of gravity is as low as 
possible. To say that the center of gravity is as low as possible is equivalent 
to saying that the potential energy of the system is as small as possible. 

The problems concerned with the determination of extreme values of 
integrals whose integrands contain unknown functions belong to the calculus 



teC. 14] INTEGRALS WITH SEVERAL VARIABLES 265 

of variations. The simplest of such problems concerns the determination 
of an unknown function y ~ y{s) for which the integral 

I = / ‘ F(r,y,y') dx (14-1) 

between two fixed points PoOoJ/o) and P\(x u yi) is a minimum. The 
function F of the variables x, y, and //' s? dy/dx is assumed to be known. 

If we imagine that the points P 0 and P t in the xy plane are joined by 
a sufficiently smooth curve y = fix), tlien the substitution of y ® f(x) 
and y f * f'(x) in the integrand of (11-1) yields the integral 1(f) whose 
value, ordinarily, depends on the choice of the curve y = f(x). We ask 
the question: What is the equation of the curve y = y(x) joining P 0 and 
Pi which makes the value of the integral (14-1) a minimum? To be certain 
that this question makes sense, it is necessary to impose some restrictions 
on the integrand in (14-1) and to specify how the curves that enter in 
competition for the minimum value of 1 are to be chosen. 

W r c shall suppose that F(v,y,y f ), viewed as a function of its arguments 
x , y, and y\ has continuous partial derivatives of the second order, and 
we assume (hat theie is a curve y ~ y(x) with continuously turning tangent 
that minimizes the integral. We then choose the competing family of 
curves as follows, bet y ~ rj(x) be any lumtion with continuous second 
derivatives which vanishes at the 
end points of the interval (.r 0 ,Xi). 

Then 

M = 0, *tri) - 0 (14-2) 

If a is a small parameter, 

g(x) * y(x) + onf(x) (14-3) 

represents a family of curves passing 
through (x {h y 0 ) and (x u yx), since 
the minimizing curve y = y(x) 
passes through these points and 
77(.r 0 ) =■ rj(xi) ~ 0. The situation 
here is that indicated in Fig. 10 The vertical deviation of a curve in the 
family (11-3) from the minimizing curve is ar}(x)] it is called the variation 
of y(x). 

Now if we substitute y and y' from (14-3) for y and y' in the integral 
(14-1), we get a function of a: 

1(a) - ( 1 F\x , y(x) + arj(r), y'(r) + ot7j'(x)] dx, (14-4) 

J *1 

*» 0, Eq. (14-3) yields y(x) = y(x), and since y - y(x) minimises 



For a 



FUNCTIONS OF SEVERAL VARIABLES 


266 


[chap. 3 


the integral, we conclude that 1(a) must have a minimum for a*0, A 
necessary condition for this is 


dl 

da a®=o 


- 0. 


(14-6) 


We can compute the derivative of 1(a) by differentiating (14-4) under the 
integral sign and get 

V(a ) « r i — F(x,Y,Y')dr t (14-6) 

Jx o da 

where we have set 


y = y(x) + ai?(x), Y’ = y'{x) + ay'(x). 
But by the rule for the differentiation of composite functions, 

dF(x,Y,Y') 


(14-7) 


da 


dF or o f or' 

0 F da d)' 1 da 


dF 

or 


dF 

v(x) + n\x), 

a } 


so that (14-6) can be written as 

OF 

I 0Y 


fT l OF 

I'M */ -M') + 

ri V 


OF 

or 


; v'M 


dr. 


Since /'( 0) = 0 by (14-5), we get, on setting a = 0 m (14-8), 
OF OF 

<y 


or or 

I ~ 

■^o L OlJ On' 


dr ~ 0, 


(14-8) 


04-9) 


because for a = 0, it is evident from (14-7) that l r = y(x), Y' = y'(r). 
The second term in the integral (14-9) can be integrated by parts to yield 


r* 1 OF OF *i rx x (i / 0F\ 

/ — v’(x)dx = -- v (x) - *(*)-(--} 

Jx o dy dy * 0 ] H dx \dy / 


0/? 


d /0F\ 


dx 


= - f\(x) — (~F)dx 
'*« dx \dy’/ 


dy'. 

since the integrated part drops out because of (14-2). Accordingly, we 
can write (14-9) as 

/•*! [" dF d ( d F\~ 

\ v(x) ) dx = 0. (14-10) 

J *t L dy dx \dy / _ 

But y(x) is an arbitrary function vanishing at the end points of the interval. 



INTEGRALS WITH SEVERAL VARIABLES 


SEC. 14} 


267 


Since the integral (14*10) must vanish for every choice of ij(x) y it is easy to 
conclude that 1 


dF d /dF\ _ 

dy dx \dy'/ 


(14-11) 


This equation is called the Euler equation . On carrying out the differentia- 
tion indicated in (14-11), we get the second-order ordinary differential 
equation 

dF d 2 F d 2 F d 2 F 


<sy 


dx dy' dy dy ' 


V - V 




(14-12) 


for the determination of the minimizing function ?/(.t). 

The general solution of (14-12) contains two arbitrary constants which 
must be chosen so that the curve y ~ y(x) passes through {x 0 ,yo) and 

It should be noted that the solution of Euler's equation (14-11) may not 
yield the minimizing curve because the condition (14-5) is necessary but 
not sufficient for a minimum. Ordinarily one must verify whether or not 
this solution yields the curve that actually minimizes the integral, but 
frequently geometrical or physical considerations enable one to tell whether 
the curve so obtained makes the integral a maximum, a minimum, or 
neither. 

Similar calculations when performed on the integral 

h V) = f ' F{x,y,y',y",. ,.,y M ) dx (14-13) 

J *0 

yield the Euler equation 

d d 2 d n 

n - r + 7 , Fy (-D- — F yin) « 0. (14-14) 

dx dr dx 


The foregoing discussion (‘an also be generalized to the problem of minimiz- 
ing the double integral 

1 ^ = IL F ( x > y ’ u ’ Ux ’ v «) dx dy ’ (14-15) 


in which the competing functions u(x } y) assume on the boundary C of 
the region R preassigned continuous values u = <p(s). If it is supposed 

1 The proof is by contradiction. Assume that the function in the brackets of (14-10) 
is not zero at some point x « £ of the interval (xo,xi). Then since it is a continuous 

d 

function, there will be a subinterval l about x * £ throughout which F v — ~ F v > has 

dx 

d 

the same sign as at x ** £. Choose ri(x) so that it has the same sign as F y — F v > in l 

dx 

and vanishes outside this subinterval. For such a choice of v(%), the integrand in (14-10) 
will be positive, and thus the integral will fail to vanish as demanded by (14-10). 



268 


FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 

that F , viewed as a function of x,y f u,u x ss du/dx, u y m du/dy, has con- 
tinuous second partial derivatives with respect to these arguments, the 
Euler equation corresponding to the integral (14-15) turns out to be 


F u 


dF Uz __ dFu 
dx dy 


(14-16) 


A special form of the integral (14-15) is of particular interest in the study 
of the Diriehlet problem, which occurs in numerous applications. 1 It is 


H «) = ff R l(u x ) 2 + («„) 2 + 2 /(*,?/)«] dx dy, (14-17) 

where f(x,y) is a known function. The substitution of F = (u x ) 2 + {u v ) 2 
+ 2 fu in (14-16) yields the Poisson equation 

V 2 u - (14-18) 

It can be shown that the solution of this equation, 2 assuming specified 
continuous values u — <p(,s) on the boundary C of the region, actually 
minimizes the integral (14-17) on the sot of all competing functions which 
take on C the same boundary values <p(s). 

Example' What is the equation of the curve y — y(x) for which the area of the surface 
of revolution got by revolving the curve about the x axis is a minimum? 

The integral to be minimized in this problem is 

I « 2tt f yds * 2w f yVl -f y' 2 dx. (14-19) 

Jx {) JjC(, 

It has the form (14-1) with 

F(x,y,y') =® 2iryV\ -(- y* 2 . ( 11 - 20 ) 

The substitution from (14-20) m the Kulei equation (14-11), after simple calculations, 


yields 

vT+ u’ 2 - d - /-£=-o, 
dxVl -f y ' 1 

or 

yy" - v'~ - l - 0. 


This second-order equation is easily solved by setting if « p, y" « p dp/dy (cf. Chap 1, 
Sec. 13). The result is 

X — f2 

y - Cl cosh > (14-21) 

ri 

so that the desired curve is a catenary. The integration constants ci and c-j in the general 
solution (14-21) must be determined so that the curve passes through given points (io,yo), 
(XhVi)* 

1 Bee Chap. 6, Sec. 12, and Chap. 7, Sec. 21. 

1 See, for example, 1. S. Sokolnikoff, “Mathematical Theory of Elasticity,” 2d ed,, 
sec. 106, McGraw-HilT Book Company, Inc., 1956. 



SEC. 15] 


INTEGRALS WITH SEVERAL VARIABLES 


269 


PROBLEMS 


1. Show that the curve of minimum length joining a pair of given points in the plane 
is a straight line. Hint: Minimize 


r 

J xn 


VI + y*dx. 


2 . Solve Prob. 1 by taking 




dr . 


3. When a bead slides from rest along any smooth curve C from the point P to a point 
Q cm C, the speed v of the bend is v **■ \ /l 2 yh, where h is the vertical distance from P to Q. 

f Q di 

: - ■ Choose P at the otigin, and show that 

Jp v 


lienee the travel time from P to Q is i 


the curve for which the travel time is a minimum is a cycloid. 

r *\ %/ 1 ~j_ ,/2 

4. Consider the integral 7 - / - - 

Jxo ' v 

the associated Euler’s equation is </ 2 E (r — r ^ c ». Discuss this solution. 
6. Obtain Euler’s equation for the mtegial 


dx, and show that the general solution of 


Hit) = J \p{r)(t /) 2 4- <lU)ip d 2f(x)y]dx. 


Special cases of this integral arise in the study of deflection of bars and strings 

16. Variational Problems with Constraints. Occasionally one seeks a 
maximum or minimum value of the integral 


/ ® f 1 FUdJd/) dx, (15-1) 

discussed in the preceding section, subject to the condition that another 
integral 

J =■ f * ('(-i',y,y') dx (15-2) 

Jx Q 

have a known constant value A physical problem of this sort has already 
l>een mentioned in Sec. 11 where it was icquired to find the shape of the 
chain which minimizes the potential energy while the length of the chain 
is given. This is one of the so-called im perimetric problems of the calculus 
of variations. 1 

It is natural to attempt to solve the problem I — min subject to the 
condition J — const by the method of Lagrange multipliers. We construct 
the integral 

I + XJ * [ 1 [F(x,y,y’) + \G(r,y,y')] dx 

J *ts 

1 Isoperimetric because the length (or the perimeter) of the curve is given. 


(15-3) 



270 FUNCTIONS OF SEVERAL VARIABLES [CHAP. 3 

and consider the free extremum of the integral (15-3). The corresponding 
Euler’s equation (14-11) is 

d(F + \G) dd(F + \G) 

= 0 (15-4) 

dy dx dy' 

and on carrying out the indicated differentiation 1 in (15-4), we get the 
second-order ordinary differential equation containing the parameter X. 
The general solution of this equation, in addition to X, will contain two 
arbitrary integration constants. The integration constants and the param- 
eter X must then be determined so that the curve y - y(x) passes through 
the given end points and satisfies the constraining condition (15-2). 

The justification of this procedure is based on an argument similar to 
that used in Sec. 14, where instead of the one-parameter family of the 
neighboring curves (14-3) one constructs a suitable two-parameter family. 2 
16. Change of Variables in Multiple Integrals. The reader will recall 

that the double integral ff R f( x >y) dA of a continuous function f(x,y) spe- 
cified in a closed two-dimensional region R of the ry plane is defined as the 
limit of the sum formed in the following way: The region R is subdivided 
into n elements of area A A t1 and the value of f(x,y) is computed at some 

n 

point (ft,^i) of the AA t ; the sum is then formed, and its 

limit is calculated when the number of elements A.4, is allowed to increase 
indefinitely in such a way that the greatest linear dimensions of the ele- 
ments tend to zero. Thus, 

r n 

I f(x,y) dA sb Wm ]£ /(£•.*».) (16-1) 

n -♦ » i 

The calculation of the limit in (16-1) is usually performed by repeated 
evaluations of two simple integrals, so that 

r rx~-h ry*~g t {x) 

I K X >V) dA — I I f(x,y) dy dr. (1G-2) 

J R Jx^a 

The limits in (16-2) are determined from the equations of the boundary 
of the region (Fig. 11). The triple integral of f(x,y,z) is defined similarly, 
by subdividing the three-dimensional region R into volume elements A r t 
and by forming the corresponding sum. Thus 

f n 

lf(x,y,z)dr= lim £/(£», At,. (16-3) 

J R n — ♦ « , __ j 

1 Compare Eq. (14-12). 

•See G. A. Bliss, “Calculus of Variations,” Carus Monograph, The Open Court Pub* 
fishing Co., LaSalle, IIL* 1925. 



SEC. 16] INTEGRALS WITH SEVERAL VARIABLES 271 

The limit of the sum in (16-3) is usually evaluated by repeated single in- 
tegrations. One can write, for example, 


f rx^h p/**h 2 (x) rz**g 2 (x,y) 

}J(.x,y,z) dr = / / I f(r,y,z) dzdydT, (16-4) 

JR ■'/rad Jy.^.h l (x) J z-^gi(x,y) 


in which the integration limits are determined from equations of the 
bounding surfaces. 

The evaluation of multiple integrals can frequently be simplified by 
making appropriate changes of the inde- 
pendent variables. Thus in dealing with 
double integrals, it may prove advanta- 
geous to replace x and // by new variables 
a and v related to x and y by the trans- 
formation 


h(i\y) 


(16-5) 



with suitable properties 

We shall suppose that the functions 
/, in (16-5) have continuous tiist partial derivatives in the region R and 
that the J acob la n of (1 6-5 ) 




du 

dv 

dx 

dx 

du 

dv 

d// 

dy 


(16-6) 


does not vanish in the region R. In this event, Eqs. (16-5) can be solved 
for x and y to yield the different iable solution 1 


X - <Pi(u,v), 
y = <P 2 (w,iO- 


(10-7) 


If u and v are assigned some fixed values, say i/o and vo t the equations 


wo = fi(*,y ) 9 


*>o = h (x,y), 

determine tw r o curves which will intersect in a point (xo,2/o), such that 


X U = <Pi(Ua t Vo) r 


y 0 = ^ 2 (wo,eo). 

1 See, for example, J. S. SokolnikofT, “Advanced Calculus/’ chap. 12, McGraw-Hill 
Book Company, Inc., New York, 1939, 



FUNCTIONS OF SEVERAL VARIABLES 


272 


[chap. 3 


Thus the pair of numbers (« 0 ,e o) determines the point (xo,yo), in the xy 
plane (Fig. 12). 



If u and v are assigned a sequence of constant values 

(»2 (UsAl), •••, (Un,Vn)j ♦ 

a network of curves will be determined that will intersect in the points 
Cri,2/l)> C**2d/2), C^lfe), * . (• Cn,Vn), 

Corresponding to any point whose rectangular coordinates are (x,y) there 
will be a pair of curves u = const and v — const, which pas* through this 
point. The totality of numbers (v,v) defines a curvilinear coordinate 
system, and the curves themselves are called the coordinate lines. 

Thus, if 

V - v?~+ 

* -l y 
v = tan -» 

x 

the family of curves u = const is a 
family of circles whereas v — const 
defines a family of radial lines. The 
curvilinear coordinate system, in 
this case, is the ordinary polar 
coordinate system (Fig. 13). 

I 11 the cartesian xy coordinates 
the element of area dA ~ dx dy is 
the area of a rectangle formed by the intersection of the coordinate 
lines x » x 0 , x = x 0 + dz, y » y Q , y = y 0 + dy, as shown in Fig, 14. In 
the curvilinear uv coordinates the element of area dA can be visualized as 




INTEGRATES WITH SEVERAL VARIABLES 


273 


SEC. 16] 

the area of the quadrilateral PiP 2 P%Pa formed by the intersection of the co- 
ordinate lines u = Uo, u » uq + du ) v = v 0 , v = v 0 + dv } shown in Fig. 15. 



The expression for the element of area dA in curvilinear coordinates ( u,v ) 
can be calculated with the aid of Eqw. (16-7), but it is somewhat simpler 
to follow the method of Sec. 2, Chap. 5 (see, m particular, Eq. (2-17) of 
that section) to show that 

dA = \J(u,v)\ dudv, (16-8) 


where 


dx 

dy 

dll 

du 

dx 

dy 

dv 

dv 


(16-9) 


is the Jacobian of the transformation (16-7). 

The double integral m (16-2) can then be evaluated in the uv coordinates 
by substituting in }{x,y) from (16-7) to obtain f\<pi(u,v), <p 2 (w,e)] ss F(u,v) 
and writing dA in the form (16-8). Thus, 

f(x,y) dA ~ / F(u,v) \J(u,v) | dudv. (16-10) 

J J v — a -In 


The limits of integration in (16-10) are determined from the equations of 
the boundary of R referred to the uv coordinates. 

Similar considerations apply to a change of variables {x,y,z) in the triple 
integral (16-4) by the transformation 

U - fi(x,y,z), 

v = h{x,y,z), 

w = h{x,y,z), 


( 16 - 11 ) 



FUNCTIONS OF SEVERAL VARIABLES 


[CHAP. 3 


with the Jacobian 


J(x,y,z) = 


If the solutions of (16-1 1) are 


du 

dv 

dw 


dx 

dx 

dx 


du 

dv 

dll) 


— 

— 

— 


dy 

dy 

dy 


du 

dv 

dw 


dz 

dz 

dz 


x — 

<P\ 

(u,v,w), 

y « 

<P2 

(u,v,w), 

z = 

*P3(U,V,W), 


(16-12) 


(16-13) 


the element of volume dr in the uvw system ran be taken as 1 
dr — \J(u,v,u>)\du dv du.', 


(10-14) 


where J(u,v,w) is the Jacobian of the transformation (16-13), so that 


J(u,v,w) 


dx 

dy 

dz 

du 

du 

du 

dx 

dy 

dz 

dv 

dv 

dv 

dx 

dy 

dz 

dw 

dw 

dw 


(16-15) 


Example 1. Let it be required to find the moment of inertia of the area of the circle 

(Lig. 16 ) 

h x 2 + y 2 ~ax~ 0 

J^ydA-pdpdS about a diameter of the circle. It is con- 

P~ a cos 0 vement to introduce the polar coordinates 

x *» p cos 0 t 

y * p gin 0, 

so that the equation of the circle becomes 
p « a cos 0. 

Calculating the determinant J gives 

_ _ j. I cos 0 sin 0 j 

^ t ~ I — p sin 0 p cos 0 I Pi 




dA ■* p dp dd. 


Pig. 16 

1 Cf. Chap. 5, Sec. 2. 


so that 




1 See Chap. 5, Sec. 1. 


Fig. 17 



276 FUNCTIONS OF SEVERAL VARIABLES [CHAP. 3 

These equations are in the form (16-13) with p » u, 6 » v f <j> * w, and it is easy to verify 
that (16-14) yields 

dr ** p 2 sin 6 dp d$ d4>. 


Consequently, the substitution from (16-16) gives 

t 1 2 

I x dr ** I / / p sin 0 cos </>p 2 sin 0 dp dd d<j> 

J R j <t> »K»0 j 6 BaO J fi «»Q 

ira 4 /16 3a 

j* 3*3 SB 

»ra 3 /6 8 


Thus, 


i6 ’ 


PROBLEMS 


1. Use cylindrical coordinates (r,6,z) defined by 

x = r coy y « / sin 0, z ~ z, 

to compute the moment of inertia of the volume of a right -circular cylinder of height h 
and radius a about its avis. Also evaluate the integral in cartesian coordinates 

2. Compute the expression for dA in forms of u and v if 

x *= u(l — v), y » uv. 

3. Compute 4 the expression for dr in terms of u, v, and w if 

x = u(l — v), y - uv, z =* uvw. 

4. Show that in the cylindrical coordinator* ol Prob 1, the element of volume dr ** 
r dr dd dz, 

6. Use the cylindrical coordinates of Piob 1 to find the volume enclosed by the circu- 
lar cylinder r — 2a cos 0, the cone z *» r, and the plane z — 0. 

6. Evaluate e~ ^ <v2; dy dx, where R is the region bounded by the circle x 2 ~f y- = 

Jr 

a z Use polar coordinates 

7. Find the area outside p — a(l -f cos 0) and inside p *= 3 a cos 6 

8. Find the cooidnuttes of the center of gravity of the area between p ~ 2 sm 0 and 
p ® 4 sin 0 , 

9. Calculate the elements of area in the uv coordinate systems which are related to 
the cartesian coordinate system ary by means ol the following equations of transformation: 

(а) x s* u 4- a, y *» v + b; 

(б) x * au, y ** bv; 

( c ) x *■ a cos a — v sin a, y ^ u sin a + t> cos a; 


where a, 6, and a are constants. Interpret your results geometrically 

10 . What are the regions of integration in the uv coordinate systems of Prob. 9 if the 
region R in the xy plane is the interior of the ellipse 


+ ^ 
a 2 ' b 2 


1? 


11 . Discuss the curvilinear coordinate system defined by the relations 


X «* U + V, 


y as u — Vj 



INTEGRALS WITH SEVERAL VARIABLES 


SEC. 17] INTEGRALS WITH SEVERAL VARIABLES 

and describe the region in the uv plane corresponding to the square x « 1, x ** 2, y 
y « 2. 

12. Discuss the curvilinear coordinate system defined by the relations 
u « z 2 ~ y 2 , v » 2xy. 


277 
- 1 , 


Sketch the curves u «*■ const and t> » const. 

13. Sliow that the attraction of a homogeneous sphere at a point exterior to the sphere 
is the same as though all of the mass of the sphere were concentrated at the center of 
the sphere. Assume the inverse-square law of force. 

14. The Newtonian potential V, due to a body T, at a point P is defined by the equa- 
tion V(P) * | wheie dm is the element of mass of the body and r is the distance 

Jt r 

from the point P to the element of mass dm. Show that the potential of a homogeneous 
spherical shell of inner radius b and outer radius a is 


V - 2ir(r(a 2 - 6*), if r < 6, 


and 



if r > a, 


whore <r is the density. 

16. Find the Newtonian potential on the axis of a homogeneous circular cylinder of 
radius a 

16. Hhow that the force of attraction of a right-circular cone upon a point at its vertex 
is 2ir<rh{] — cos a), where h is the altitude of the cone and 2 a is the angle at the vertex 

17. Show that the force of attraction of a homogeneous right-circular cylinder upon a 
point on its axis is 


2ira\h + Vl{* + a 2 - VOi'+h)* + a s J, 


where h is altitude, a is radius, and It is the distance from the point to one base of the 
cylinder. 


17. Surface Integrals. A surface is usually defined as a locus of points 
determined by the equation 

^ = /(.rj/), (17-1) 

where f(x,y) is a continuous function specified in some region of the xy 
plane. Tins definition, however, is too broad to permit one to formulate 
a meaningful concept of the surface area. Since most surfaces encountered 
in applications are two-sided and piecewise smooth, we confine our con- 
siderations to such surfaces only. 

The surface defined by (17-1) is called smooth if it has continuous partial 
derivatives dz/dx and dz/dy at each of its points. This implies that a 
smooth surface has a continuously turning tangent plane and hence a 
well-defined normal at each of its points. 1 


1 We recall that the equation of the tangent plane to (17-1) at a point P(zo,lM)) is 

“(£)/- x(,) + GX (!/ - vo) ’ 

> that the direction of the normal at P is determined by the ratios ( — ) : ( ~ ) : — 1 

\dx/p \dy/j[> 


(cf. Sec. 10, Chap. 4). 



278 FUNCTIONS OF SEVERAL VARIABLES |CHAP. 

The surface is said to he pieceurisc smooth if it can be subdivided b, 
smooth curves into a finite number of pieces, each of which is smooth 
Thus, the surface of a cube is piecewise smooth. 

The surface is two-sided when it is possible to paint it with two differen 
colors to distinguish the sides. 1 If two oppositely directed normals PI 
and PN' (Fig. 18) are drawn at a point P of a smooth two-sided surfac 
and P is allowed to move along any path that does not cross the edge c 
the surface, the direction of PN can never be brought into coincidenc 
with PN\ 



It is intuitively clear that a small element of a smooth surface is nearl 
flat, so that a neighborhood of any point on it is well approximated by 
portion of the tangent plane. This observation suggests a procedure fc 
constructing a meaningful definition of the area of a smooth surface. 

Thus, let S' be a smooth portion of the surface S bounded by a close 
curve C (Fig. 19). We shall .suppose that S' is such that every line paralh 
to some coordinate axis (say the z axis) outs S' in just one point. If th 
projection C' of C on the xy plane encloses the region R , we can subdivici 
R into n small subregions AR t by the families of straight lines parallel t 
the x and y axes. The planes through these lines, normal to the region 1 
cut from S' small regions AS' of areas Aa a . Let. AA t be the area of A R 
The projection of Aa t on the xy plane is, approximately, 

A.4 1 = cos 7 * A (7 2 ,, 

where cos cos &, and cos y » are the direction cosines of the norm; 

1 At first glance, it may appear that all surfaces are two-sided, but this is not the cae 
A simple example of one-sided surface, whose boundary is a closed curve, is given 
Sec. 6, Chap. 5. 



SBC, 17 ] INTEGRALS WITH SEVERAL VARIABLES 

N to 8 at a point (x,,^,*,) of &$[. Since 1 


and 


cos a t : cos & : cos y t ~ 


(-M-V 1 

\dx/ % \dy/ l 


cos 2 a x + cos 2 ft + cos 2 y % = 1 


we have 


cos y t = 


-i 

dr y/(dz/dx)l~ f ( dz/dijif + 1 


279 


Using the' positive 1 value for cos y t) which amounts to the choice of the posi* 
tive direction of N, wc can write 


A <r, == sec y t A A, 



+ 1 *A t . 


The surface area of S' can then be approximated by the sum 


P,M + 0‘, +> ^ 


and we define the area a of S' by the integral 



(17-2) 


The integral (17-2) can be evaluated by repeated integrations to yield, 
for example, 


a 





+ 1 dij dx. 


By considering the projections IV and It" of S' on the other coordinate 
planes, we deduce similar lormulas: 



To obtain the surface area of a piecewise smooth surface we need merely 
to add the areas of its smooth pieces. 

The surface integral of a continuous function <p(x,y,z) specified on the 
surface S' is defined as follows: Let S' be subdivided into subregions AS[ 


See the first footnote in this section. 



280 FUNCTIONS OF SEVERAL VARIABLES 

of areas Act* and form the sum 


[chap. 3 


2 <pfa>Vit z i) ( 17 - 3 ) 

t— l 

where {x^y^Zi) is some point in AS[. The limit of the sum (17-3) as n — ► *> 
in such a way that the greatest linear dimensions of the A S' % tend to zero 
is the surface integral of <p(x,y,z) over S'. It is denoted by the symbol 

f s , <p(*,y> z ) da. (17-4) 

The integral (17-4) can be evaluated by repeated integrations. Thus, if 


da = sec y dA 




+ 1 dx dy, 


then <p(x,y,z) da =* 1L <e{x,yj{x,y)) ^(— ) + + I dx dy 


where z = f(x,y) is the equation of S' and R is the projection of S' on the 



dz ~x 

dx Vo 2 — x 1 — y 1 


xy plane. 

We shall consider surface integrals 
in somewhat greater detail in Chap. 
5. 

Example 1. Find the surface of the 
sphere x 2 4* V 1 4* & m a 2 cut off by the 
cylinder x 2 — ox -f y 2 « 0 (Fig. 20). 

From symmetry it is clear that it will 
suffice to determine the surface in the first 
octant and multiply the result by 4. Now, 


dz ^ -y 
dy \/ a 2 — x 2 — y 2 


Thus, the integral becomes 


«r 



1 dy dx 



V °* = S adydx 

\/ a* — x* — y* 


BBC. 17] INTEGRALS WITH SEVERAL VARIABLES 281 

It is simpler to evaluate this integral by transforming to cylindrical coordinates. The 
equation of the cylinder becomes r « a cos 0, and that of the sphere 


Thus, 


z » Vo 2 — x 2 — y 1 »» Vo 2 — r*. 


or 



ar dr dd 



Example 2. Find the z coordinate of the center of gravity of one octant of the surface 
of the sphere x 2 -f* V 2 + & m a 2 . Now, 


/,- rr%/©’+©'+> 

CBC 

[ da 
Js' 


4 ir a 2 
~8~ 


dj dy 


a dz dy 

_ a 

^2 ~ 2 * 


PROBLEMS 

1. Find, by the method of Sec. 17, the area of the surface of the sphere x 2 -+* l/ 2 4* 
z 1 ** a 2 Unit lies in the first octant. 

2. Find the surface of the sphere z 2 -f- ?/ 2 + z 2 =» a 2 cut off by the cylinder x 2 az -f- 
/ -* 0. 

3. Find the volume bounded by the cylinder and the sphere of Prob. 2. 

4. Find the surface of the cylinder x 2 -j- y 2 — a 2 cut off by the cylinder y 1 -j- z 2 *** a 2 . 
6. Find the coordinates of the center of gravity of the portion of the surface of the 

sphere cut off by the right-circular cone whose vertex is at the center of the sphere. 

6. If a sphere is inscribed in a right-circular cylinder, then the surfaces of the sphere 
and the cylinder intercepted by a pair of planes perpendicular to the axis of the cylinder 
are equal in area. Prove it. 




CHAPTER 4 


ALGEBRA AND GEOMETRY OF VECTORS. 


MATRICES 




Fundamental Operations 

1. Scalars, Vectors, and Equality 287 

2. Addition, Subtraction, and Multiplication by Scalars 288 

3. Base Vectors 291 

4. The Dot Product 293 

5. The Cross Product 294 

6. Continued Products 297 

7. Differentiation 299 

Applications 

8. Mechanics and Dynamics 202 

9. Lines and Planes 306 

10. Normal Lines and Tangent Planes 309 

11. Frenet’s Formulas 311 

Linear Vector Spaces and Matrices 

12. Spaces of Higher Dimensions 316 

13. The Dimensionality of Space. Linear Vector Spaces 317 

14. Cartesian Reference Frames 321 

15. Summation Convention. Cramer's Rule 324 

16. Matrices 327 

17. Linear Transformations 332 

18. Transformation of Base Vectors 337 

19. Orthogonal Transformations 340 

20. The Diagonal ization of Matrices 343 

21. Real Symmetric Matrices and Quadratic Forms 347 

22. Solution of Systems of Linear Equations 350 


285 




It is desirable to treat directed quantities like force or velocity (which 
are independent of coordinate* systems) witfiout reference to a set of co- 
ordinate axes Such a coordinatc-fm? treat ment is made possible )>y the 
analytical shorthand known us vector analysis. The trajectory of a 
part id**, the dynamics of rigid bodies, and the theory of fluid flow are 
readily studied by vector methods, as are also such topics as the geometry 
of nines and surfaces. Introduction of coordinates yields a correspond- 
ence between vectors and sets of numbers, and this correspondence permits 
the use of vector methods in the study of linear equations. Such a study 
leads to the concept of mainx , which has proved fruitful in a variety of 
licit Is, ranging from circuit analysis to quantum theory. 


FUNDAMENTAL OPERATIONS 

1* Scalars, Vectors, and Equality. Some quantities appearing in the 
study of physical phenomena can be completely specified by their magni- 
tude alone Thus, the mass of a body can be described by the number 
of grams, the temperature by degrees on some scale, the volume by the 
numbci of cubic units, and m> on. A quantity that (after a suitable choice 
of units) can be completely characterized by a single 
number is called a scalar There are also quantities, 
called vectors, that require for their complete charac- 
terization the specification of direction as well as mag- 
nitude An example of a vector quantity is the dis- 
placement of translation of a particle. If a particle is 
displaced from a position P to a new position P' (Fig 
1), then the change in position can be represented 
graphically by the directed line segment PP* whose 
length equals the amount of the displacement and whose direction is from 
P to P\ Similarly, a force of magnitude K dynes can be represented by 
a line segment whose length is K units and whose direction coincides with 
that of the force. 



287 



288 


ALGEBRA AND GEOMETRY OP VECTORS, MATRICES [CHAP, 4 

The initial point P of a directed line segment representing a vector is 
called the origin , and the representation as an arrow suggests that the 
terminal point be called the head of the vector. In many problems the 
location of the origin for any given vector is immaterial, and in such 
problems two vectors are regarded as equal if they have the same length 
and the same direction. Such vectors, which need not coincide to be equal, 
are termed free vectors . In mechanics, it is sometimes convenient to 
specify vectors by giving the line of action as well as the length and di- 
rection. Equality of these so-called sliding vectors means that the lengths, 
directions, and lines of action coincide. Again, in the treatment of space 
curves and trajectories one is led to specify the origin of the vector 
as well as its length and direction. Such vectors are termed bound 
vectors . 

To distinguish vectors from scalars, boldface type is used for vectors in 
this book. The length (or magnitude) of the vector A is denoted by | A | : 

| A | = length of A. (1-1) 


Equality is denoted by the usual symbol: A — B. For the most part this 
chapter deals with free vectors, and hence “A = B” means that A and B 
have the same length and direction. 

2. Addition, Subtraction, and Multiplication by Scalars. If a particle 
is displaced from its initial position P to P\ so that PP' = A, and sub- 
sequently it is displaced to a position P", so that P'P" = B, then the 
displacement from the original position P to the final position P" can be 

accomplished by the single displacement PP" = C. Thus, it is logical 
to write 



Fig. 2 


A + B - C 

as the definition of vector addition (Fig. 2). In 
words, if the initial point of the vector B is placed in 
coincidence with the terminal point of the vector A, 
then the vector C, which joins the initial point of 
A with the terminal point of B, is called the sum 
of A and B and is denoted by A + B =* C. This 
is the familiar parallelogram law of addition used in 
physics, and its extension to three or more vectors 


is obvious. The symbol + behaves like the + of elementary algebra, in 


that 


A + B = B + A, commutative law 
A + (B + C) = (A + B) + C, associative law. 


(2-1) 



BBC. 2] FUNDAMENTAL OPERATIONS 289 

A proof is implicit in Figs. 3 and 4. The associative law enables us to omit 
parentheses, writing A + B + C for A + (B + C). 




It is desirable to give meaning to expressions like 5A, the product of a 
scalar and a vector. In agreement with the meaning of multiplication 
familiar from arithmetic, one defines 

5A = A + A + A + A + A (2-2) 

(Fig 5) and similarly in other cases. By a natural extension of this reason- 
ing, /A is defined as a vector whose length is |/| | A| and whose direction is *hat 


A 


-* 


5A 

Fig. 5 


of A if t is positive but opposite to that of A if t is negative. One defines At 
by the equation At _ lA (2 . 3 ) 

It follows that 1A — A and also 

s(tA) = (st) A, associative law, 

( s + t ) A = sA + /A, distributive law, (2-4) 

t(A + B) = tA + tB, distributive law. 

A vector of zero length is denoted by 0 and termed the zero vector. To 
introduce the idea of subtraction, one defines —A as the solution of the 
equation A + X = 0. Evidently, —A is a vector equal in length to A 
but of opposite direction, so that —A = ( — 1)A. As in elementary 
algebra, B — A is used as an abbreviation for B + (—A). 

Since the laws governing the addition of vectors and multiplication of 
vectors by scalars are identical with those met in ordinary algebra, one is 
justified in using the familiar rules of algebra to solve linear equations involv- 
ing vectors. 



290 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [('HAP. 4 

Example 1. The point P in Fig. 6 divides the segment AB in the ratio m:n. Express 
R in lei ms of the vectors A, B and the scalars m, n. 


A 



— > 

The vector X * AB satisfies A + X 
hence, solving for X, 

X 


B by the definition of vector addition, and 
B - A. (2-5) 


(This exemplifies the so-called “head-minus- tail” rule for vector subtraction.) The vec- 

— y 

tor /IP is m/(m -f n) times X, by the hypothesis and by the definition of multiplication 

> 

by scalars. Since R — A -fr- AP we have, finally, 


R 


A + - X -(B - A) 
rn 4- n 


nA + mB 
m -f n 


Exa?nplc 2. Prove that the medians of every triangle intersect at a point two-thirds 
of the way from each vertex to the opposite side. 

Let two sides of tin' t mingle* Ik* specified by veetois A and B, as in Pig 7, so that the 
third side is B — A (cf Example 1). The vector median to the side B — A is 1 ^ the diag- 



onal of the parallelogram on A, B; hence this median is (A -f B)/2 (compare the special 
case m « n of Example 1). If the point M in Fig. 7 is two-thirds of the way from the 
vertex 0 to the side B — A, along this median, then 


SKC. 3] 


FUNDAMENTAL OPERATIONS 


291 


2 A + B A + B 

°M -“ 8 — 


The vector median to the side A is A/2 — B, again by the head-minus- tail rule If N 

is at a point two-thirds of the way toward the side A on this median, then 




Comparison with (2-(5) shows that the two points M, N coincide. That tiie third median 
also lias the required behavior follows by interchanging the roles of A and B. 


PROBLEMS 

1. Sketch a vectoi A of length 1 5 in , paialicl to the lower edge of your paper and 
having an arrow on its light-hand end Sketch a second vector B ol length 1 in., making 
ari angle of 30° with A Now sketch 2A, 3B, A + B, A — B, 2A — 3B, (A -f B)/2 

2. (live a condition on three veetoi.s A, B, C which ensures that they can form a tri- 
angle (Generalize to n vectors A, B, C, . . , L 

3. (liapbiealiv and algebraically, show how to find twm vectors A and B if their sum 
S and difference D arc known 

4. Sketch three vectors A, B, C issuing from a common point. On your figure show 
the vectors A — C, B — A, C — B, and thus illustrate the algebraic identity (A—C)*r 
(B - A) 1 (C - B) = 0 

5. (a) Wnte down a vector ot unit length which has the same direction as a given 
nonzero vector A {!>) Using t lie result (a), wiitc down a vector bisecting the angle formed 
by two nonzero vectors A, B issuing from a common point 

6 Show that a line fiom a ve? tex of a paiallelogram to the mid-point of a nonadjacent 
salt' trisects a diagonal. 

3. Base Vectors. Any vector A lying in the plane of two noncollinear 
vectors a and b can be resolved into so- 
called components directed along a and b. 

Tins resolution is accomplished by con- 
structing the parallelogram whose sides are 
parallel to a and b (Fig S). Then one can 
write 

A = xa + yh, 

where x and y are the appropriate scalars. 

If three noncoplanur vectors a, b, and c are given, then any vector V 
can be expressed uniquely as 

V = xsl + yb + zc } (3-1) 

where V is the diagonal of the parallelepiped whose edges are xa, yb } and 
zc (Fig. 9). The vectors a, b, and c are called the base vectors , and the 
scalars x, y, and z the measure numbers. 

An important set of base vectors, denoted by i, }, and k, consists of unit 
vectors directed along the positive directions of the x, y, and z axes, re- 



a xol 


Fkj. 8 



292 


ALGEBRA AND GEOMETRY OP VECTORS* MATRICES [CHAP* 4 

spectively (Fig. 10). It is assumed that the system of axes is a right- 
handed system; that is, a right-hand screw directed along the positive 
z axis advances in the positive direction when it is rotated from the positive 




x axis toward the positive y axis through the smaller (90°) angle. Because 
i, j, k are mutually orthogonal, the representation 

A a xi + y] + zk 
yields the important formula 

| A |* «** + ** + ** (3-2) 

by use of Pythagoras's theorem (Fig. 11). 



Example: If A = 1 + 2j + 3k and B =» ~j -f- 4k, compute the length of 2A — B. 
Since 2A - B « 2i -f- 5j -f 2k, we have 

|2A - B| - (2® + 5 2 «f 2*)« - V33. 


SEC. 4] 


FUNDAMENTAL OPERATIONS 


PROBLEMS 

1. (a) In the form oi 4* 4* ck write down two vectors of length 5 parallel to the y 

axis, (b) If A * 14* 2j + 3k, B * i + j 4* k, C » i - k, compute A + B, (A + B) *f 
C, B + C, and A -f (B 4- C). What law does this illustrate? (c) In (6), find 5A, — 2A, 
the sum of these vectors, and the vector 3A. What law does this illustrate? (d) Also 
find 3A, 3B, the sum of these vectors, and the vector 3(A 4- B). (c) In (6), a certain 
vector D is such that A, B, D can be placed head against tail to form a triangle. What 
is the z component of D? 

2. Sketch the triangle with vertices at the heads of i 4* j 4* k, 2j + k, and 2i -f j, 
and make the sides into vectors with head against tail. Find the vectors forming the 
sides of the triangle, and verify that; the sum of these vectors is zero. 

8. Draw a figure illustrating the inequality |A + B| < | A | 4* |B|, and by combin- 
ing this with (3-2), deduce an algebraic inequality. Can you give a purely algebraic 
proof? 

4. (a) I .»et A, B, C, ... be vectors from the center to the vertices of a regular decagon 
(ten-sided polygon). By choosing a suitable basis i, j and using symmetry, show that 
the sum A ■+• B + C + •*• is zero ( b ) By picking another basis i', j', with i' making an 

angle 6 with A, deduce the identity cos 9 4* cos (9 4* ir/5) 4* cos (9 4* 2r/5) 4 h 

cos (9 4" Or/5) ** 0. 

4. The Dot Product. The dot product 1 of two vectors is defined to be 
the product of their lengths by the cosine of the angle between them. In 
symbols, 

A-B - | A] | B [ cos (A,B), (4-1) 

where (A,B) is the angle from A to B. Thus A-B is a scalar , not a vector. 
Geometrically, 

A-B = | A j X (projection of B on A) 

«= | B j X (projection of A on B). (4-2) 

Evidently (A,B) can be measured in several ways. However, since cos d 
» cos (— 6 ) = cos (2r — 6 ), these different measures all yield the same 


value for A*B. The fact that cos 0 =*• cos ( — 6) also yields 

A*B = B-A, commutative law, (4-3) 

and one easily verifies the additional properties 

(£A)-B = f(A-B), associative law, (4-4) 

A*(B + C) = A-B + A-C, distributive law. (4-5) 

For proof of (4-5) use (4-1) to transform (4-5) into 

| A | | B 4* C | cos yp » | A | | B | cos 4> + I A | | C | cos 9, (4-6) 

where the angles are defined in Fig. 12. Now (4-6) follows from 

| B 4- C | cos $ « | B | cos 4> 4- I C | cos 9 (4?-7) 

1 The terms scalar product and inner product are often used. 



294 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

and (4-7) is evident from Fig. 12* when the vectors are coplanar and the angles are in 
the first quadrant. In view of (4-2) the property amounts merely to the assertion that 
projections are additive, and the extension to arbitrary angles is not difficult, 

For the mutually orthogonal unit 
vectors i, j, k introduced in Sec, 3 
we have, by inspection of (4-1), 

i-i = j*j = k-k = 1, 

(4-8) 

i*j - J*k - i*k - 0. 

Fin. 12 Hence, expanding the product by 

(4-4) and (4-5), we get 

(ri 4- ?yj + 2 k)-(.rji + /yj + z{k) = xx x -f yy x + zz x . (4-9) 



By (4-1) and (4-0) the dot product gives a simple way to find the angle 
between two vectors and, in particular, to decide when two vectors are 
perpendicular. Indeed, if we agree 1 to regard the zero vector as perpen- 
dicular to every vector, then from (4-1) 

A-B = 0 if, and only if, A JL B. (4-10) 


The case in which B is parallel to A is also worthy of note, 
when B = A we have . . , . l2 


In particular 
( 1 - 11 ) 


Example. Compute the cosine of the angle between A and B if A ** 
B « — i zk, and find a value of z for which A ± B. 

We have A-B «-1-f0+2z-22-l and hence, by (4-1), 


cos (A,B) 


2 2 - 1 2z -J 

JaFTbT " 


The result is zero, and hence tin* vectors are perpendicular, when z = 


+ j + 2k, 


PROBLEMS 

X. Given A = i 4- 2j -j- 3k, B — — i 4- 2j 4 k, C = 2i 4~ j- («) Find the dot prod- 
uct of 3i 4* 2j 4- k with ea^h of these vectors (h) Find A-B, A-C, B 4 C, A-(B 4- CL 
What law is illustrated? (c) Find 2A and (2A)-B Compare A-B as found in (/>). (d) 
Find the angle between A and B (c) Find the projection of A on C (/) Find a scalai s 
such that A 4- tfB is perpendicular to A. (g) Find a vector of form ix 4~ j y 4“ k which 
is perpendicular both to A and to B 

2. (a) Show' that i -p j 4- k, i — k, and i — 2j 4- k are mutually orthogonal. (6) 
Choose x f y, and z so that i 4- j 4" 2k, — i 4~ 2k, and 2i 4- *rj 4- yk are mutually 
orthogonal. 

3. (a) If A-B = A-C for some A 5^ 0, is it necessary that B » C? Illustrate your 
answer by an example. (6) If A-B = A-C for every A, is it necessary that B * C? 

5. The Cross Product. Besides the multiplication just considered there 
fs a second kind of multiplication, which yields a product known as the 



FUNDAMENTAL OPERATIONS 


295 


EEC. 5] 

vector product or cross prodvct. The cross product of A and B, denoted 
by A x B, is a vector C which is normal to the plane of A and B and is 
so directed that the vectors A, B, C form a right-handed system. The 
length of C is the product of the length of A by the length of B by the 
sine of the smaller angle between them: 

|A x B| ~ j A | | B | sin (A,B). (5-1) 

The expression (5-1) represents the area of the parallelogram having A, B 
as adjacent edges (Fig. 13). The student is warned, incidentally, that 




(5-1) does not give A x B; it gives the length |A x B' only. 

Since rotation from B to A is opposite to that from A to B, we have 

A xB = — B x A, (5-2) 

so that the commutative law does not hold for vector products. On the 
other hand it is the case that 

(/A) x B = /(A x B), associative lav, (5-3) 

A x (B + C) ~ A x B + A x C, distributive law. (5-4) 

The proof of equation (5-3) is trivial, and (5-4) i.-. readily established if we note that 
A X V is obtained from the arbitrary vector V by performing the following three opera- 
tions 0 , illustrated in Kig 1 1 

Oil Project V on the plant* perpendicular to A to obtain a vector \ L J_ A of mag- 
nitude | VI sin (A,Vj. 

Og: Multiply Vi by |A| to obtain V 2 1 A of magnitude |A| |V| sin (A r V). 

O 3 " Rotate V 2 about A through 00° to obtain V.-s = A x V. 

It is easily chocked that each of these operators is distributive; that is, 0 t (B -f C) ** 
O t B -f OjC for all vectors B and C Hence the composite operator O 3 O 2 O 1 is distributive; 
namely, 

0 3 O 2 Oi(B 4~ C) * 0,i0v(0iB 4- OjC), since Oi is distributive, 

= OziO^Oi B 4~ O 2 O 1 C), since O 2 is distributive, 

— O 3 O 2 O 1 B 4 O 3 O 2 O 1 C, since O 3 is distributive. 

Because 0/H ) iV ** A X V for every vector V, the latter equation yields (5-4). 



296 ALGEBRA. AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

The definitions of vector product and of i, j, k lead to 

ixi=*jxj*=kxk = 0, ixj=— jxi-»k, 

j x k * — k xj = i, k x i «= -i x k - j. (5-5) 

If A and B are given by their components as 

A = xi + y] + zk, B = Zji + j/jj + z^ 

then expansion by means of (5-3) and (5-4) and simplification by means of 
(5-5) yield 


A x B = i (yz, - zyi) + j(x x z - zz,) + k(xyi - yx x ) 
which may be written as a determinant 1 


A x B 


i j k 



= i 

y z 


X z 

+ k 

x y 

X y z 

- j 




Vi zi 


Xi Zi 


2-1 2/1 


xi y i z i 


(5-0) 


Example . Find a vector perpendicular to i -f 2k and i -f j — k, and fmd the area of 
the triangle with these two vectors as adjacent sides. 

Both questions are settled by calculating the cross product. We have, from (5-6), 


(i + 2k) X (i + j - k) 


i j k 

1 0 2 

1 1 -1 





0 

1 


- ~2i + 3j + k. 


This vector is perpendicular to the given vectors The area of the triangle is half the 
area of the parallelogram: 

Area A*=3^l~2i + 3j+k| «* %\/l4. 


PROBLEMS 

1, Given A - i + 2j + k, B - 3i + 2j, C - ~i + j + 3k. (a) Find A x B, A x C, 
Ax B+Ax C, B + C, and A X (B + C). What law is illustrated? ( b ) Find a vector 
perpendicular to B and C, and verify your answer by use of the dot product, (c) If A, 
B, C have their origins at a common point, find a vector perpendicular to the plane in 
which their heads lie. (d) Find the area of the triangle formed by the heads in (c). 

2. Show that the cross product for each two of the following vectors is parallel to 
the third: i 4" j k, i — k, i — 2j 4* k. What does this indicate about the vectors? 

, 1 The reader unfamiliar with second- or third-order determinants is referred to Ap- 
pendix A. 



SEC. 6] FUNDAMENTAL OPERATIONS 297 

3 . Give an example of three unequal vectors such that the cross product of any two is 
perpendicular to the third. 

1 If A X B * 0 and A«B « 0, is it necessary that A * 0 or B » 0? 

6. In refraction at the plane interface of two homogeneous media let A, B, C be unit 
vectors, respectively along the incident, reflected, and refracted rays, and let N be the 
unit normal to the interface, (a) Show that the law of reflection is equivalent to A X II 
■BXN. ( b ) Show that the law of refraction is equivalent to r^A X N «* » 2 C X N, 
where n\ and are the indices of refraction. 

6. Continued Products. With the two multiplications previously de- 
fined, we can form the products (A*B)C, A*(B x C) and A x (B x C); 
some of the other possible combina- 
tions, however, have no meaning. 

For example, (A^B) x C is mean- 
ingless because the two factors in a 
cross product must both be vectors. 

The first product, (A*B)C, denotes 
simply the product of the scalar 
A'B with the vector C and may be 
dismissed without further comment. Fig. 15 

By definition of dot product, the 

second expression, A* (B x C), called the scalar triple product , has the value 

A*(B x C) - | A | cos 0 1 B x C|, (6-1) 

where 0 is the angle between A and B xC. Since B x C is perpendicular 
to the face of the parallelepiped containing B and C (Fig. 15), and since 
|B x C| is the area of this face, (6-1) shows that A • (B x C ) represents the 
signed volume of the parallelepiped having A, B, C as adjacent edges. More- 
over, we have the formula 

A X A y A z A — i A x 4' $A y -f* kA|, 

A-(B x C) * B x B y B z , B = \B X + j B v + k B t , (6-2) 

C x C y C z Q = iC x 4~ jCy + k C M> 

as will now be seen. The expression (5-0) yields 

i j k 

B x C - B x B y B t » iP + jQ + k R, (6^3) 

C x C y C z 

say, where P, Q, R are certain second-order determinants. Taking the 
dot product of iA x + \A V + kA* with (6-3) leads to 

A*(B x C) *= A X P "4* A y Q *4- A z R f 

which is the expansion of the determinant (6-2) on elements of the first row. 




298 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

Since interchanging two rows of a determinant merely changes its sign, 
(6-2) yields the useful relations 

A*(B xC)« B*(C x A) = C* (A x B) 

- — B* (A x C) - — A* (C xB)= -C(B x A). (6-4) 

These results as to magnitude are evident from the volume interpretation, 
though further discussion is needed to establish the algebraic sign in this 
way. Because of (6-4) it is customary to write 

A-(B xC)= A*B x C - (ABC). (6-5) 

To evaluate the vector triple product A x (B x C), let i be a unit vector 
parallel to B and j a unit vector perpendicular to i in the plane of B and C. 
Thus 

B = B x \, C = Cxi + Cyj, A = A x i + A v ] + A z k, (6-6) 

where k is a unit vector perpendicular to i and j. so oriented that the three 
form a right-handed system. Since B x C = B x C y k by (6-6) and (5-6), 
we have 


A x (B x C) = -A x B x C y } + A y B x C v x 

* (A X C X + A y C y )B x i - A x B x {CA + C y j) 

~ B(A'C) - C(A-B). (6-7) 

Example' Establish the identity 

! A'C BC I 
(A x B) • (C x D) = j A .p b-D I' 

The expression is the scalar triple piodurt of A x B, C, and D. Interchanging the dot 
and cross, as we may by (0-4), wo obtain 

(A X B)*C X D * (A X B) x C D - |(A C)B - (B C)AJ*D 

- (A-C)(B-D) - (B-Cj(A-D), (6-9) 

since (A X B) x C « - C X (A x B) - (A-C) B - (B-C) A by (6-7). 

PROBLEMS 

1* Verify (6-2), (6-7), and (6-8) by direct calculation for the special case A ** i 4* j, 
B ** ~i + 2k, C « j + 2k, D - i + j + k. 

2. (a) In Prob. 1 find the volume of the parallelepiped having A, B, and C as adjacent 
edges. (6) Find x such that the vectors 2i -f j — 2k, i -f- j -f- 3k, and ri -f j are coplanar. 
B%nt. A certain parallelepiped must have zero volume, (c) State a simple necessary 
and sufficient condition that three arbitrary vectors A, B, C be coplanar. (d) Evaluate 
(AAB) and (ABA), where A, B are arbitrary. 

$. By (6-7) show that A x (B x C) + B x (C X A) + C X (A X B) « 0. 

4. Show that (B X C) X (C X A) « C(ABC), and deduce 

(A x B)-(B x C) x (C X A) * (ABC) 2 . 



SEC. 7] FUNDAMENTAL OPERATIONS 299 

6. The vectors A, B, C issue from a common point and have their heads in a plane. 
Show that (A X B) + (B X C) ~f~ (C X A) is perpendicular to this plane. 

7. Differentiation. If for each value of a scalar t a vector R(/) is defined, 
we say that R is a vector function of t. In a particular problem t may 
denote the time and R the position vector of a moving point relative to 
some origin ( ). As in the calculus of scalars, we say that R (t) is a contin uous 
vector function of t at t = t 0 provided that 

lim R (t) = Rfo). (7-1) 

t - 1 0 

The precise meaning of (7-1) is that |R(f) - R(/<>) | becomes as small as 
desired whenever t is sufficiently near t {) . 

The cartesian components of the vector R (t) are functions of t, so that 
one may write 

R(0 “ hr (t) + MO + kz(t). (7-2) 

It follows from (7-1) that the functions x(t), y(t), z(t) are continuous if, 
and only if, R(/) is continuous. 

We define the derivative of R(0 with respect to t by the formula 

(7R R (t + AO - R(0 

-- - lim (7-3) 

dt At -+ o A t 


The substitution of (7-2) in the definition (7-3) leads immediately to the 
result that R is differentiable if, and only if, x } y , z are, and in that case 


(JR 

It 


dx d // d* 
i r + j T + k-: 


dt 


dt 


dt 


(7-4) 


As in scalar calculus we shall write R'(<) for dR/dt y R"(0 for d 2 R/dt 2 , and 
so on. 

Products involving vectors are differentiated by the familiar rules of 
elementary calculus, and the proof of these rules also involves only familiar 
ideas. For example, the formula 


dt 


dB dk 

(A x B) - A x — d x B 

V dt dt 


(7-5) 


follows from 


A(A x B) = (A + AA) x (B + AB) — A x B 

= A x AB + AA x B + AA x AB 

when we divide by At and let At — ► 0. Of course, the order of the factors 
in (7-5) must be preserved, since the cross product is not commutative. 



300 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

A geometric interpretation of the derivative may be obtained as follows: 
Let the vector R (t) be regarded as a bound vector with its origin at the 
origin of coordinates. The head of R then traces out a space curve as t 
varies (see Fig. 16). The vector 

AR = R(f + AO - R(0 (7*6) 



is directed along a secant of the curve, AR/A i is parallel to this secant, 
and hence lim (AR/AO is tangent. Thus, the vector R'(0 is tangent to the 
Space curve R = R(<) whenever R'(0 exists and R'(0 9* 0. 

To interpret the magnitude jR'(0|, let s be the length of the curve 
from the fixed point given by t *= ^ to the variable point given by t. 
Assuming R'(0 ^ 0, we have AR 0 for small At > 0, and hence 


As 

At 


As | AR| 
I AR! At 


As 

IARI 


AR 

At 


(7-7) 


Since |ARj is the length of the chord, and since the ratio As/|AR| of 
arc to chord 1 tends to 1, Eq. (7*7) gives 


ds 

dt 


dR 

~dt 


(7-8) 


when At —» 0. Thus, the vector R'(0 has magnitude |R'| « ds/dt where s 

1 We assume that s increases with t; otherwise a minus sign is needed. The fact that 
late) /(chord) -* 1 follows from the familiar interpretation of arc as limit of lengths of 
inscribed polygons. It is also possible to take (arc) /(chord) I as one of the defining 
properties of arc and proceed, as in the text, to obtain the formula (7-9). 



FUNDAMENTAL OPERATIONS 


BBC. 7] 


301 


is the arc length along the curve. If R '{t) is continuous, the arc is 
explicitly by 


* - £ |R'(0 1 dt = £ Viy? + (i V '? + W dt. 


given 

(7-9) 


Introduction of s as parameter instead of t facilitates the study of space 
curves (see Sec. 10). 


In two dimensions the interpretation of R'(/) given here agrees with the results of 
elementary calculus. Let a smooth curve C be represented parametrically by x ** x(t), 
V * 2/(0, so that the slope is given by 


Slope - ~ 
ax 


dy/dt 

dx/dt 


x' 


(7-10) 


for x' 7 ^ 0. If the same curve is described in the form R « ix + jy, we have R' ■* ix' -f- 
j y', and hence the slope of the vector R' is y'/x' In view of (7-10), the fact that R' w 
tangent to the curve agrees with the fact that dy/dx is ike slope of the curve. The formula 
ds/dt |R'| is also familiar; it states that 

*.V5VTW ->/©■+ (*)’. 

which liecomes ds 2 *» dx 2 -f dy 2 when squared and multiplied by (dt) 2 . 

Physically, one may regard t as time, so that the head of the bound 
vector R(f) gives the position of a moving particle at time t . Since the 
velocity is defined to be V = R '(t), the foregoing result means that the 
velocity vector is tangent to the trajectory and has magnitude equal to the speed 
ds/dt i with which the particle is moving . 


Example 1. The position of a particle at time t is determined by the bound vector 
R(0 » if + j^ 3 4- k sin t. 

Find a vector tangent to the orbit at time and find the speed of the particle at time 
t «* 0 . 

We have R'(0 « i + 3j t 2 4* k cos t, which is the required tangent vector. At t *» 0 
the velocity is R'(0) « i + k, and hence the speed is ds/dt *= |R'(0) | *= s/2. 

Example 2. If a differentiable vector R(l) has constant length, show that R' is per- 
pendicular to R, and interpret geometrically. 

From R*R ® const, differentiation yields R*R' -f R -R = 0, whence R'*R » 0. 
Geometrically, if R is a bound vector of constant length, its head traces out a curve lying 
on a sphere. The tangent to the curve is tangent to the sphere, hence perpendicular 
to the radius vector. Thus, R' A R. 


PROBLEMS 

1. If R(0 « I2( + fit 2 + k f 3 , (a) find the derivative R'(0- (&) At the point (2,3,1) 
find a tangent to the space curve which is traced out by the head of R when R is regarded 
as a bound vector. Hint : The point (2,3,1) corresponds to t « 1. (c) If R(<) is a bound 



302 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

vector giving the position of a moving particle at time t, find the velocity and speed of 
this particle at time I *» 1. 

2. (a) Differentiate the vector R(0 » it + j sin t 4* k cos t, compute |R'(0 1, and sim- 
plify. (b) If R(£) is a bound vector, find the length of the curve traced out by the head 
of R as t varies from t « 0 to t » 2. 

3. By writing A«B in component form and differentiating, deduce (A-B)' **> A'*B + 
A*B'. 

4. If Ro and A are constant, find a vector tangent to the curve described by the bound 
vector R « Ro + At. 

3. If R(/) is a bound vector giving the position of a moving particle at time t , the 
acceleration is defined to be A * R "(/). Show that A is constant if R(0 ** Ro 4* Ri* 4- 
R 2 i 2 , where Ro, Ri, and Rj are constant; vectors. Is the converse true? 

6. Show that (ABC)' - (A'BC) + (AB'C) 4* (ABC '), when A, B, C are differentiable, 
and write out in determinant form. 

7. If R = A -f /(OB, where A and B are constant and / is twice differentiable, then 

R r X R" - 0. 

APPLICATIONS 

8. Mechanics and Dynamics. The work IT done by a constant force F 
producing a displacement S in the direction of F is |F| |Sj. More gener- 
ally, if F makes an angle 6 with S, the work is |Fj |S| cos 0, and hence 

W - F-S. (8-J) 

Because of this equation the dot 
product plays a central role in cer- 
tain branches of mechanics. 

To illustrate the application of 
cross products, let the vector ft rep- 
resent the angular velocity of a 
rotating body; that is, let & be a 
vector whose magnitude is the angu- 
lar speed in radians per second and 
whose direction is parallel to the axis 
of rotation. The positive sense of 
ft is chosen as that in which a right- 
handed screw would advance if the 
screw were rotated in the same di- 
O rection as the body. Let R be a vee- 

Fig. 17 tor locating any point P of the body 

relative to some point 0 on the axis 
of rotation. Tt is required to find the instantaneous velocity V of the 
point P If the distance of P from the axis of rotation is a r then by Fig. 17 

|V| - |ft |a - |ft| | R | sin (R,Q). 

Moreover, V is normal to the plane of R and ft and is so directed that 




SEC. 8] APPLICATIONS 303 

fl, R, and V form a right-handed system. Hence, 

V = 0 x R. (8-2) 

The result is independent of the origin O, for if a new origin O x is chosen and P is 
specified by a vector Ri from Oi, then 

Ri - R + S, 

where S is parallel to (see P'ig. 17). Hence Q x S = O, and therefore 

QxR]«flx(R + S)*QxR + ftxs*flxR. 

Another example from dynamics illustrates the compactness of vector 
notation. Let 0 be a fixed point in a rigid body, and let a force F be applied 
at a point R of the body, which is located by the bound vector R whose 
origin is at (). The force F establishes a torque or moment T which tends 
to rotate the body about an axis that passes through O and is normal to 
the plane of R and F. The magnitude of T is given by 

I T | - ! R | |F|sin(R,F). 

In addition, R, F, and T form a right-handed system, so that 

T - R x F. (8-3) 

That Ihe choice of 0 is immaterial follows as in the discussion of (8-2). 
Similarly one shows that F may slide along its line of action without 
affecting the result ; that is, F may be regarded as a sliding vector. 

To illustrate the use of (8-3), we obtain a formula for the so-called center of mass of a 
system of mass points. The force on a point of mass m in a gravitational field is given 
by mF, where m is the mass of the point and F is a vector specifying the strength of the 
field at the point in question. We assume a uniform field, so that F is independent of 
position. From (8-3) 

(R - P) x mF (8-4) 

represents the moment about the point 1 P of the gravitational force on a point of mass 
m at It. If there are n points of masses m i, m 2 , . . ., m n located by the vectors R 1} R 2 , 
. . R n , respectively, the total moment about the point P due to all of them is 

2(R» - P) X m t F (8-5) 

It is desired to find a single mass point such that its moment (84) reproduces the 
total moment (8-5) for all choices of F and P. Equating the moments (84) and (8-5) 
leads to 

[mP - 2m*P - mR + R»] X F « 0, (8-6) 

after rearrangement. Since F is arbitrary in (8-6), the factor in brackets must vanish, 
so that 

P(m — 2/rq) = mR — Sm^R*. (8-7) 

1 The vectors R, P, and R, are bound position vectors with a common origin for the 
points /?, P, and R t , respectively. 



304 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

The fact that P is arbitrary in (8-7) now gives 

to » Xim t to R « Sm,R t . (8-8) 

Conversely, (&-8) ensures the validity of (8-7) and hence of (8-6) independently of F and P. 

This discussion was carried out by equating moments only. Equation (8-8) shows, 
however, that the total gravitational force is also preserved, since the mass of the point 
equals the total mass of the collection. 

The point R with position vector 

m{Ri + m 2 R.2 + m n R n 

R “ T “ T (&-9) 

mi + m 2 -i 1 - m n 

determined by (8-8) is called the center of mass. Evidently the collection 
of points, regarded as a rigid body, wpuld balance about the point R as 
pivot, for the moment (8-4) is zero when P = R, and hence the moment 
(8-5) also vanishes. 

Still another example of the use of vectors in mechanics is given by 
Newton's laws. Relative to an origin 0, which is regarded as fixed, let 
the position of a particle at time t be specified by the bound vector R (f). 
The velocity vector V is dR/dt , as indicated in Sec. 7, and the momentum 
vector is defined by 

dR 

M « mV = m — » (8-10) 

dt 


where m is the mass of the particle at time t. In this notation Newton's 
second law of motion takes the simple form 


F - 


dM 

dt ’ 


mi) 


where F is the force on the particle at time t. 


F - 



If m is constant the result is 
( 8 - 12 ) 


We shall use (8-10) and (8-12) to derive some interesting properties of the center of 
mass. Suppose given n particles with masses m,- and positions denoted by R t (i ** 1, 
2, ...» n), where each rm is independent of t. The total momentum of the system 


satisfies 


Xrrii 


dRi d 
dt dt 


2m, R* 


d ZrruRi 
dt m 



(8-13) 


where to «* 2m, is the total mass and where R locates the center of mass [(8-9)]. Thus, 
the total momentum of the system equals that of a single particle which has mass m and moves 
with the same velocity as the center of mass of the system. 

If (8-13) is differentiated with respect to t, there results 


2F, 


TO 


d 2 R 
dt * 


When we let F* be the force on the t'th particle and use (8-12). Since internal forces can- 
ed in pairs by Newton's law of equal and opposite reaction, the sum 2F, represents the 



APPLICATIONS 


SBC. 8] 


305 


total external foroe acting on the system. Hence the center of mast has the same accelera- 
tion ae a particle of mass m acted on by a force equal to the sum of the external forces acting 
on the system. 

Example 1. Parallel farces F, — F of equal magnitude but opposite direction constitute 
a couple. Find the total moment, and show that it is the same about every point 
Let R be a vector from a given point 0 to a point P on the line of action of F, and Ri 
to a point Pi on the line of action of — F (Fig. 18). The total torque is 


RxF+RjX (-F) 


(R-Ri)xF - (PiP) x F. 


Since this is independent of 0, the result 
follows. Notice that F and — F must be 
regarded as sliding vectors (Sec. 1) rather 
than free vectors, since the line of action 
is fixed. 

Example 2. A system of forces F* acting 
at various points R% of a rigid body is such 
that 2F« «■ 0. If the total torque about 
one point is zero, then the total torque 
about every point is zero. 

From 2(Ro — R») X F* «* 0, say, we arc 
to deduce 2(R — R*) x F» * 0. The two 
equations may be written 

He x (SFJ 

R x (2F») 



Fig. 18 

2R t x F t , (8-14) 

2Rt x Ft. (8-15) 


Equation (8-14) gives 2R» X F, * 0, since 2F t = 0, and (8-15) follows. 

Example 3. The moment of the momentum vector M about a point is called the 
angular momentum of the particle about that point. According to the principle of angu- 
lar momentum, the rate of increase of angular momentum about a point equals the re- 
sultant torque about that point. Show that this principle is equivalent to Newton's 
law, F * dM/dt 

If A is the angular momentum about the oi igin, then A » R X M, where R gives the 
position of the point. Thus 


dk 

It 


R x 

R x 


dM 

dt 

dM 

dt 



x M 


+ V x (mV) 


R x 


dM 

dt 


The principle of angular momentum dk/dt 


R 


dM 

dt' 


R x F is therefore equivalent to 
R x F. (8-16) 


If this holds for every choice of origin, that is, for every R, then necessarily dM/dt » F. 
Conversely, if dM/dt - F, then (8-16) holds for every R. 


PROBLEMS 

1, Given A •» i 4- 2j 4- k, B - 1 - k, C - 2i + j, with A, B having their origins at 
a common point, (a) find the work done by a force A in a displacement B. (h) Find 



306 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

the work done in a displacement from the head of A to the head of B under a force C. 
(c) Find the work done in the displacement A subject to simultaneous forces B and C. 

2. In Frob. 1: (a) Find the torque about the origin of A due to a force C through the 
head of A. (6) Find the torque about the head of A due to a force C acting through the 
head of B. 

3. In Prob. 1 : (a) If the figure formed by A and B rotates about A with angular veloc- 
ity ft, find the velocity of the head of B. (b) Find the velocity of the head of A if the 
figure formed by A and B rotates with angular velocity ft about an axis parallel to C 
through the head of B. 

4. Two coordinate systems have a common origin at all times, but the second has a 
vectorial angular velocity ft relative to the first. Show that Vj *» V 2 -f (ft X R), where 
V x and V 2 are the velocity vectors in the first and second systems of a point whose posi- 
tion vector is R in the first system. 

5. Show that the torque due to two couples is the sum of the torques. 

G. Three points labeled 1, 2, 3 have masses 1, 2, 3 and positions 2i -f j 2k, i — k, 3j, 
respectively, (a) Find the center of mass, (b) Find the total mass 2, 1, and their cen- 
ter of mass. From this obtain, again, the center of mass for all three. 

7. The vectors A, B, C, D, E give the positions of the vertices of a regular pentagon 
as referred to an origin not necessarily in its plane. Show that their resultant is equal 
to 5R, where R gives the position of the center Hint * Place a unit mass at each ver- 
tex, and find the center of mass in two ways. 

8. (a) Show that F*V represents the rate at which work is done on a particle moving 
with velocity V under a force F. ( b ) When the mass is constant, show that 

(d/dt)(m |V| 2 /2) - F-V, 

so that the rate of increase of kinetic energy equals the rate at which work is done on 
the particle. 

9. Lines and Planes. If R is a bound vector with its origin at the origin 
of coordinates, then the direction numbers x , y , z are the same as the 
coordinates of the head of R, and one may speak indifferently of “the 
point R“ or “the vector R.” This correspondence between vectors and 
points enables us to use vectors in geometry. Here we consider the ge- 
ometry of lines and planes, which is especially simple; the following sections 
are concerned with general curves and surfaces. 

Suppose we have given a plane through the point R<> and perpendicular 
to the constant vector A. If the point R is in the plane, then R — R 0 
is perpendicular to A, and conversely (Fig. 19). Hence the equation of 
the plane is 

(R — Ro)*A ~ 0. (9-1) 


If D is the distance from the point Ri to the plane, then 


D = | R x — Ro | | cos 0 1 


1 Rr -RollAMcosfll ^ KR l ~ RoVAI 

~ ~ |A| ~ TaT 


(9-2) 


where 6 is the angle between A and Ri — Ro. 

Next, consider a line through the point Ro and parallel to a constant 
vector A. If the point R is on this line, then the vector R — Ro is parallel 



J3ML 9] APPLICATIONS 307 

to A, and conversely (Fig. 20). Hence, the equation of the line is 


(R — Ro) x A = 0. (9-3) 

If I) is the perpendicular distance from the point Ri to this line, then 



Fm 19 Fig 20 


Tn (9-3) the fact that R — Ro is parallel to A may also be expressed 
by writing 

R - Ro - At, 

where t is a scalar. Thus we obtain the equation of the straight line in a 
parametric form, 

R - Ro + At, -~oo < t < oc, (9-5) 

which is often more useful than (9-3). It is left to the student to deduce 
the cartesian equation by setting 

Ro = a 0 i + Vo) + -ok, A =* ai + b) + ck 

in (9-5) and equating components. Eliminating t yields the symmetric 
form 

x “Z° - y ~~ lJo _ 

a b c 

which may also be found from (9-3). 

Example 1. Show that every equation of form 

ax -f by 4- cz + d » 0 a, b, c, d const 
represents a plane with A » ai 4- hj + ck as normal, and conversely. 


(9-7) 



308 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

If R is a general point and Ro a fixed point on the locus (9-7), then writing (9*7) in 
vector form yields 

R«A -fd ** 0, Ro-A -f d « 0. 

Subtracting these equations we obtain (9-1), which shows that the locus is a plane. On 
the other hand, (9-1) itself has the form (9-7), with d * — Ro-A, and hence the converse 
is also true. 

Example 2. Find the equation of a line which passes through the point i — j and is 
parallel to the two planes x -f- y « 3, 2x 4- y -f 3* » 4. 

The respective normals to the planes are i + j and 2i 4* j 4* 3Jk, and hence the line 
of intersection of the planes is parallel to the cross product: 

(i -f j) x (2i + j + 3k) * 31 - 3j - k. 

Since this vector is parallel to both planes, it gives the direction of the required line, 
and hence the equation is 

R « i - j + (3i - 3j - k )t, ~co < t < oo. 

PROBLEMS 

1. (a) Find a vector normal to the plane x + 2y 4* 3i * 1. (b) Find the angle be- 

tween this plane and the plane af + y + *+ 2*0. (c) What is the distance from the 
point 3i 4* 2j -f k to the plane in (a)? ( d ) Show that the points i and — j -f k lie in 
the plane in (a), (e) Find a vector lying m the plane in (a). Hint- Subtract the vectors 

of (d). (/) Verify that the vector of (c) is normal to the normal found in (a). 

2. (a) Find a vector parallel to the line R — i + k -f (i -f 2j 4- 3k) t. (b) If R »■ ix 4* 

jy 4- kz in (a), find x, y, and z in terms of t. ( c ) In (a), find the distance from the point 
i 4- 2j 4- 3k to the given line. ( d ) Show that the line ( a ) intersects the line R * 2k 4- 
(Si -f 2 j -j- k)«. Hint' Equate the two expressions for R, and consider each component. 
It will be found that all three equations are satisfied by s ~ f =» (<0 Find the inter- 

section point in (d). (/) Find the point where the line in (a) intersects the plane 2x — 
y 4- Sz «* 4. Hint: Substitute the result of (b) into the equation of the plane, find t, 
then find R. 

3. (a) Find the equation of the line common to the two pianos x 4- 2j/ 4* 4* m 1 t 

x 4- V ** 3 in the form R * Ro 4- AL Hint: Let z » t, and solve for x and y in terms 
of t. (b) Find a vector parallel to the intersection of the planes by use of the cross prod- 
uct as in Example 2. (c) Verify that your answers to (a) and ( b ) are consistent, (d) 

Find the equation of all planes perpendicular to both planes, (e) Write the equation 
of the line which is parallel to both planes and passes through the point — 3i 4- k. 

4. (a) In terms of t, find the square of the distance from the point i 4- 2j 4“ 3k to a 
general point on the line R 3i 4- 2j 4" k 4- (i 4- j 4- k)f. (b) By differentiating, find 
the t for which the distance is minimum and the minimum value, (c) Check by the 
distance formula. 

fi. In the form (9-5) obtain the equation of a line perpendicular to the plane i + v + 
3z ■* 0 at the origin. At what point does this line intersect the plane y ** 3z -j- 1? 

6* If the lines R ■■ Ro 4- At and R » Ri 4“ Bt are not parallel, then the perpendicular 
distance between them is 

|(R! -Ro) -AX B| 

|AXB| 

Hint: By a suitable figure show that the distance is the length of the projection of 
Ri — Ro on the common perpendicular to the two lines. 



me. 10] APPLICATIONS 809 

10. Normal Lines and Tangent Planes. If a curve C:x « z(t), y * y(t), 
z = z(t) lies on a surface which has the equation 

u(x,y,z) * c, (10-1) 

where c is constant, then 

4^(0,y(0,z(0] s c (10-2) 


identically in t. At a fixed point R 0 = ix 0 + jy 0 + k z 0 (Fig. 21) we dif- 
ferentiate (10-2) by the chain rule (Chap. 3, Sec. 4) to obtain 


du dx du dy du dz 
dx dt dy dt dz dt 


(10-3) 



This may be written as 

n*R'(0 - 0, (10-4) 

where R(0 « ix(0 + j y(t) + k z(t) and where 
du du du 

n = i f j h k — at (x 0l y 0 ,z 0 ). (10-5) 

dx dy dz 

Since R'(0 is tangent to the curve C Ly Sec. 7, it follows from (10-4) 
that n is normal to the curve C. And since this is true for every choice of C, 
the vector n must be normal to the surface. The tangent plane is the 
plane perpendicular to n at R 0 , and hence its equation is n* (R — Ro) = 0 
by Sec. 9. 

The assumptions which underlie the foregoing result are clear from the derivation. 
We assume u differentiable (so that the chain rule holds), and we assume that not all the 
partial derivatives are zero (otherwise n «= 0, and n does not determine a direction). 
The analysis shows, then, that n is perpendicular to every differentiable curve R «* R(t) 
which passe s through the point Ro and lies in the surface . It is this property that enables 
us to consider n as “normal to the surface.” 



310 


AtrGEBRA AND GEOMETRY OF VECTORS MATRICES [CHAP. 4 

To illustrate the use of (10-5) we find a normal vector and tangent 
plane for the ellipsoid ^ + 2 y 2 + 3 z 2 = 12 

at the point (1,2, -1) Since u = x 2 + 2 y 2 + 3 z 2 , the partial derivatives 
are 2x, 4 y, and 62 Evaluating these at (1,2, — 1) and substituting in 
( 10 - 5 ) give the normal vector 

n == 2i + 8j — Ok 

The tangent plane is perpendicular to n and contains the point (1,2,— 1) 
Hence its equation is 

x -f* 4y — 3z ~ 12, 

as the reader can verify 

Introduction of the tangent plane leads to a simple interpretation of the differential 
(Chap 3, Sec 3) If the equation of a surface is given in the form z *= f(x,y ), then 

f(x y y) - z * 0 

and hence (10-1) holds with u{x y y z) « f(x,y) — z By (10-5) a normal is 

n = l-- + j^~k (10-6) 

Ox dy 

so that the tangent plane has the equation 1 

(x - J 0 ) % + ( V - Vo) “ * * — *o, (10-7) 

Ox 0y 

where df/dx and Of/Oy are evaluated at (To, Vo) If we set x — xq » At, y — Vo “ A y, 
and z — 20 « Az in (10-7) (Fig 22), there results 

Of Of 

At d Ay = A z 

Ox du 



Fig 22 

1 The values t, y, z in (10-7) refer to the tangent plane and must not be confused with 

the values x, y, z on the surface z * f(x t y) 



APPLICATIONS 


311 


SEC. 11] 

The left-hand side is simply the differential df, and hence the differential for the surface 
z » /Or, 2 /) equals the increment for the tangent plane. The definition of differentiability 
given in Chap. 3, Sec. 3, now lias a simple intuitive meaning; namely, f{x,y) is differen- 
tiable if, and only if, the surface z ®» f{x,y) u well approximated by its tangent plane . 


PROBLEMS 


1. By use of (10-5) find a vector normal lo the plane ax -f by -f cz -f d » 0. Com- 
pare Sec. 9, Example I . 

2. At the point (2,1,3) on the surface xyz — jc 2 -f 2 find (a) a normal vector, (6) an 
equation for the tangent plane, (r) an equation for the normal line. 

3. Show that the surfaces xyz ■» 1 and x 2 -} ip — 2 zr ** 0 intersect at right angles at 
the point (1,1,1); that is, the tangent planes are perpendicular 

4 . The two surfaces x 2 «f y 2 +2^6 and 2x l 3 <r 4 r = 9 intersect at (1,1,2). 

Find the angle between the tangent planes at this point 

6. In Prob 4 find a vector tangent to the curve m which the surfaces intersect. Hint • 
The required vector is perpendicular to both normals 


11. Frenet’s Formulas. It was .shown in See. 7 that the vector R'(/) 
is tangent to the spare curve R = R(/) and has length |R'| — ds/dt, 
where s is the are along the curve If (he parameter itself is equal to the 
are, so that t = s and 


then ds/dt = 1. 

T 


R - R( s ), 

In this ease the vector 

_ dR 
ds 


(ii-i) 


(11-2) 


is a tangent vector of urn l length. 

From T*T — 1 we deduce that 
(IT /ds is perpendicular to T (Sec. 7, 

Example 2). Hence we may write 

(IT 

- - xN, x > 0, (11-3) 

(Is 

where N is a unit vector perpendi- 
cular to T and where x is a scalar 
multiplier. The vector N defined 
by (1 1-3) is called the principal nor- 
mal, and the scalar x is called the 
curvature . The plant 1 of T and N 

is termed the osculating plane . We define x =* 0 for a straight line. 

If we introduce a third unit vector B defined by B - T x N, then the 
system T, N, B forms a right-handed set of orthogonal unit vectors, analo- 
gous to the vectors i, j, k introduced previously. By Fig. 23, 

N xB = T, B x T = N, TxN-B. 



/c 


(11-4) 



312 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

The vector B is called the binormal; the figure formed by T, B, W is some- 
times referred to as the trihedral associated with the curve. 

Differentiating the relation B « T x N and using (11-3) give 

B , = TxN' + rxN = TxN , + (*N) x N « T x N', 

and hence B' is perpendicular to T. It is also perpendicular to B, since 
B*B m 1, and therefore B' is parallel to N: 

dB 

— - rH. (11-5) 

as 

The scalar multiple r in (11-5) is called the torsion; it measures the rate 
at which the curve twists out of its osculating plane. We define r = 0 for 
a straight line. 

To evaluate dN /ds, recall that N = B x T. Hence 

N' « B xT' + B' xT = xB xN + rN x T (11-6) 
by (11-3) and (11-5). When we use (1 1-4), Eq. (11-6) reduces to 

dN 

— =-*T-rB. (11-7) 

ds 

Equations (11-3), (11-5), and (11-7) are known as the Frenet-Serret for- 
mulas; they are of fundamental importance in the theory of space curves. 

By equating the lengths of the two vectors in (11-3) and recalling |N | = 1 we obtain 

«-|T'|-|R"|, '-£• (11-8) 

To get a similar formula for r we differentiate (11-3), obtaining 

T" - *N' + x'N « *(-*T - rB) + *'N (11-9) 

by (11-7). Hence, by (11-3), (11-9), and (11-4), 

r X T" *xNX(-, 5 T- xrB + x'N) « * 3 B - * 2 rT, 


since 5 X N » 0. Taking the dot product with T yields 

T-T' x r « ~* 2 r. (11-10) 

If we solve (11-10) for r, express x 2 in terms of R by (11-8), and express T in terms of 
R by (11-2), there results 

R'*R" X R'" 

r « — — (11-11) 


which is the desired formula. When R 
give, respectively, 


ir(s) + j y(s) + k*(«), Eqs. (11-8) and (11-11) 



- (*")* + iv"? + (*")\ 


( 11 - 12 ) 



SEC. 11 ] APPLICATIONS 313 

It can be shown that x(s) and r(«) determine the curve completely, apart from it* position 
in space. 1 


Since a smooth curve can always be expressed in terms of its arc as 
parameter, the foregoing theory suffers no loss of generality by assuming 
t — 8. In many physical problems, however, it is more fruitful to take 
the time t as parameter, and this possibility is now to be examined. 

Let R = R(/) give the position of a moving particle at time t f so that 
the velocity is V = R'(f). With v — ds/dt we have 2 


dR dR ds 

V - - - = Tv, 

dt ds dt 


( 11 - 13 ) 


upon using (11-2). Since (11-3) gives 


we get 



d T 

d T ds 




z=s — 

= xNr, 


dt 

ds dt 

dV 

dv 

dT 

dv 

- 

- T — 

+ V ~“ 

= T — 

dt 

dt 

dt 

dt 


+ *r 2 N 


(u-ii) 


upon differentiating (11*13) Hence the acceleration vector A = (IV /dt lies 
in the osculating plain , its tangt ntial component has magnitude equal to the 
linear acccU ration dv/dl , and its normal component has magnitude xv 2 . 
This is a far-reaching generalization of the familiar results 


Atanp’c'ntinl — 


Anomial 

r 


for uniform motion in a circle of radius r. 


Taking the cross product oi (11-13) and (11-11) with V replaced by R' we obtain 
R' x R" - T X N = *c 3 B. 


Hence, the direction of the binormal is given bv R' X R" 

even w hen the parameter is 

l rather than s Since B is a unit vector, v\e have 



R' x R" 

, d 


B ~ -! 

|R X R" | 

“ dt 

(11-15) 

and similarly, the unit vector T is obtained from 



m R ' 

d 


" IR'f 

It 

(11-16) 

Knowing B and T we find N from 



N - B x T. 


(11-17) 

'Se®, for example, L. P. Eisenhart, “An Introduction 

to Differential Geometry,” 


sec. 6, pp. 25-27, Princeton University Press, Princeton, N.J., 1910. 

1 In agreement with the results of Sec. 7, Eq. (1 1-13) expresses the fact that the veloc- 
ity is tangent to the orbit and has magnitude equal to the speed. 



314 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES |CHAP. 4 

These formulas enable us to compute the trihedral when the curve is given with an 
arbitrary parameter i provided ds/dt > 0. 

Example 1 . Find the equation of the osculating plane at t * 1 for the curve R =* t\ 4- 
2( 2 j + t»k. 

Differentiation gives 

R'(l) » i 4~ 4j 4- 3k, R"(l) - 4j -f 6k 
Hence by (11-15) the binormal B is parallel to 

(i + 4j + 3k) X (4 j -f 6k) - 12i - 6j -f 4k. 

The osculating plane is normal to B ami contains the point 

R(l) - i -f 2j f k. 

Hence its equation is G.r — 3 y 4~ 2z = 2, as the reader can verify. 

Example 2 A curve is a plane curve if, and only if, the torsion is zero 
If the curve is a plane curve, not a straight line, then the oscillating plane is well 
defined and is the plane of the curve. Hence B is constant, and the toision vanishes bv 
(11*5). Suppose, conversely, that the torsion is zero Then B is constant by (11*5), 
ft&d therefore, using (11-2), 

■^(B-R) = B-® - B-T = 0. 
ds da 

This gives B*R * const, which is the equation of a plane. 

Example 3. Consider the circular helix (Fig. 21) with equation 

R «* ia cos 0 -f ja sin 0 4“ kptf, a, p positive const. (11 -IS) 

Here, the parametric equations are 


x «* a cos ft, 
y — a bin ft, 


z — pft. 

By (11-8) x ~ JR" |, wheie primes denote differentiation with respect to the arc param- 
eter #. From (11-18) 

dR — — i a sin $ dft + j a cos 0 dft -f k p dft , 

so that 

ds 2 a* <£R*dR = (a 2 sm 2 0 T cos 2 ft -f p 2 ){dft) 2 = ( a 2 4- p 2 ) do 2 
and therefore dft/ds ** 1/Va 2 4- p ? *■ h, say. It follows that 
dR dR dft 


ds dO ds 


( — ia sm 0 4- yi cos 0 -j- k p)h. 


d 2 R 
ds 2 

d 3 R 
ds 3 


d dR dft , „ 

» ( — jo eos 0 ~ sm ft)h \ 

dft ds ds ' 

dd 2 Rdft „ . . _ 

T" *7? ~r ** mu 6 — jtt COH 0M ♦ 
dft ds* ds 



s&c , 11} 


APPLICATIONS 


315 



Fio 21 


On making use of loimula (11 -S) wo find 

= (R' ‘R") a* (a 2 sitr 0 -f a 2 cos 2 0)/> 4 = « 2 /i 4 , 


that 

Amnding to (11-12) tho toision is 


a 

*> , «> * 
<r -f /F 


- a sin 0 a (*ob 0 p 
-a iosfl ~ a sin 0 0 

a bin 0 —a cos 0 0 


a 2 -j- p 2 


If p a= 0, we got a circle of radius a by inspection of (11-18) In this case r * 0 because 
the curve is a plane curve and * * 1 /a because the radius is alw ays equal to the constant 
a. The behavior as p ~+ qo may be discussed similarly. 


PROBLEMS 

1. Given the curve 1 t(t) = i(t 2 — 1) -J- 2/j -f- ( t 2 + l)k. (a) Find a unit tangent at 
t « —1. (b) Find the equation of tho normal plane at this point, (r) Find the length 
of the curve from t « 0 to t * 1. 

2. (a) If R(/) in Prob 1 represents an orbit, find the velocity and acceleration at time t. 
(b) By use of (a) and Eq. (11-13), find the speed t> at time t (c) By use of (a), (6), and 
(1 1-14) find the curvature x and the principal normal N at time t. 



316 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP, 4 

3. If the components of R(0 are second-degree polynomials in t, then R *» R(0 is a 
plane curve, (a) Prove this by use of (13-12) and Example 2. (b) Find the equation of 
the plane. 

4 . Show that (a) the tangents to a helix make a fixed angle with the axis of the helix, 
(b) the principal normal is perpendicular to the axis of the helix. 

fi. (a) Given a particle moving according to the law R(0 « it -f j t 2 , find a unit tangent 
and a unit normal to the orbit at t » 1, (6) Find the cartesian components of V and A 
at t «* 1. (c) By use of the dot product and (a), find the tangential components V t and 

A t of V and A at t « 1 . (d) Find ds/dt as jR'(0 |, and from this find dh/dt 2 . Com- 
pare (r). 

6. In Prob. 5, (a) show that V n , the normal component of V, is zero, and find that of 
A at t * 1 by use of the dot product and Prob. 5a. (b) By (a) and A n **■ * j V | 2 find the 

curvature of the orbit at t «* 1. (c) Show that the cartesian equation of the orbit is 

pi 5 , and compute the curvature by x » y f 7(1 + j/' 2 )**. Compare (b), (d) Explain 
how to find A n in terms of A*, A„, and A<, and use this to check some of your work. 


LINEAR VECTOR SPACES AND MATRICES 

12. Spaces of Higher Dimensions. There is nothing mysterious about 
the idea of spaces whose dimensionality is greater than three. In locating 
objects in the familiar three-dimensional space of our physical intuition, 
we have found it convenient to introduce a coordinate system and to 
specify the location of any point in the object by means of three numbers 
termed the coordinates of the point. Thus, if a cartesian system of axes 
is introduced, we can associate with each point P an ordered triple of 
labels ( x,y,z ). 

In dealing with the state of gas determined by the pressure p, volume 
v } and temperature T, it is often useful to visualize the triples of values 
ip>v f T) as coordinates of points in three-dimensional space, but such a 
visualization fails when the number of variables characterizing the gas- 
state exceeds three. Thus, the state of gas may (and generally does) 
depend not only on the pressure, volume, and temperature, but also on 
the time t. Although a quadruple of values (p,v,T,t) cannot be represented 
as a point in a fixed coordinate system in the three-dimensional space, the 
geometric visualization is of much lesser importance than the analytic 
apparatus developed for coping with the geometric problems. This ap- 
paratus (analytic geometry and vector analysis) makes use of the tools 
of algebra and analysis which involve operations on ordered sets of quanti- 
ties such as (p^VjTf) or (xi,x 2 ,. . which are valid regardless of the 
number of variables appearing in the set. 

The habits of using the language associated with geometric thinking 
are so strong, however, that it is natural to continue speaking figuratively 
of r quadruple of numbers (p,v,T,t) as representing a point in four-di- 
mensional space and more generally refer to an ordered set of n values 
. . . ,z n ) as a point in n-dimensional apace . The values x h x 2) . . , , x n 



LINEAR VECTOR SPACES AND MATRICES 


317 


SEC. 13J 

may be of quite diverse sorts; the first three, for example, may be as- 
sociated with cartesian coordinates of some point M in three-dimensional 
physical space, x 4 may represent the magnitude of electric charge located 
at M, Xs may stand for the time of observation, and so on. But whatever 
meaning we choose to attach to the individual values x %y we can speak 
of the n-tuple (x\,X 2 , . . ,,x n ) as representing a point P in n-dimensional 
space. 

In three-dimensional space we found it useful to associate with every 

pair of points Pi and P 2 an entity P\P 2 which we called a vector a, and 
we have developed a set of rules for operations with vectors which form 
the basis for the algebra and calculus of vectors. 

Although in the initial formulation of these rules we have been guided 
by geometric considerations, we have distilled out geometry by giving a 
set of algebraic laws (2-1), (2-3), (2-4), (4-3), (4-4), and (4-5) which govern 
operations with vectors 

We can continue using the suggestive language of three-dimensional 
vector analysis and say that every pair of points P\ } P 2 in n-dimensional 
space determines a vector a. We further stipulate that in devising the 
rules for operating on such vectors we adopt the set of algebraic laws (2-1), 
(2-3), (2-4), (4-3), (4-4), (4-5), which contain no reference to the dimen- 
sionality of space, and we define the vector 0 by the relation a + 0 <* 0 + 
a = a for every vector a. 

The dimensionality of space, we recall, entered only when we made use 
of these laws in those calculations which involved the representations of 
vectors by components in special coordviaU systems . Thus in Sec. 3 we 
considered a vector in the plane determined by a pair of noneollinear 
vectors and introduced the notion of base vectors and the so-called com- 
ponents of the vector along the base vectors. We also saw that a vector 
in three-dimensional space can be represented uniquely in terms of its 
components in the directions of three noneoplanar base vectors. These 
remarks suggest that the dimensionality of space is in some way connected 
with the number of base vectors needed to represent a given vector by 
components. In providing a generalization of the representation of vectors 
by components in spaces of higher dimensions, we need the notion of 
linear dependence of a set of vectors which we develop next. 

13. The Dimensionality of Space. Linear Vector Spaces. The concept 
of linear dependence of a set of vectors a*, a 2 , . . a n is intimately con- 
nected with the idea of dimensionality of space. 

Definition. A set of n vectors ai, a 2 , . . a n is linearly dependent if 
there exists a set of numbers a\ y a 2) . . . , a*, not all of which are zero , such that 


aqai + a 2 & 2 H — * + <*na n ==* 0. 


(13-1) 



318 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 


If no such numbers exist , the vectors a i5 . . ., a n are said to be linearly 
independent } 

To get at the geometric meaning of this definition consider two vectors 

a and b which are like or oppositely 


directed (Fig. 25) Then we can find a 
number k ^ 0 such that 


Fro. 25 


b - Aa. (13-2) 


We can write this equation in symmetric form by setting k — —a/0, so 
that (13-2) reads 

<*a + 0b = 0. (13-3) 


Since neither a nor 0 is zero, it follows from our definition of linear depend- 
ence that two collmear \ectors are always linearly dependent Inasmuch 
as every vector b directed along a can be represented! in the form (13-2), 
formula (13-2) serves to define a imc-dimcnsional linear rector space . We 
observe that every two vectors in such a space are linearly dependent. 
If we consider two noncollinear vectors a and b (Fig. 2C>), then every 
vector c in their plane can be represented in the form 


c — Aqa -j- Aqb 


(13-4) 



Fig 20 Fig. 27 


by a suitable choice of the constants A*i and A 2 . Equation (13-4) can be 
written as 

aa + 0b + 7 c - 0, (13-5) 

in which not all constants a , 0 f y are zero. Formula (13-4) determines 
every vector c in the plane of a and b, and it thus defines a two-dimensional 
linear vector space , while formula (13-5) ensures that every three vectors 
in the two-dimensional space are linearly dependent. 

If we take three noncoplanar vectors a, b, c (Fig. 27), we can represent 
every vector d in the form 

d *= Aqa + A 2 b + A 3 c, (13-6) 

1 Of. the definition of linear dependence of a set of functions in Sec. 21, Chap. 1, 



SEC. 13J LINEAR VECTOR SPACES AND MATRICES 319 

from which follows the relation 

aa + 0b + 7 C + 5d = 0, (13-7) 

in which a, 0, % 5 are not ail zero. 

Equation (13-7) states that in a three-dimensional linear vector space 
defined by (13-0), four vectors are invariably linearly dependent. 

The foregoing discussion indicates a relationship between the dimen- 
sionality of a vector space with the number of linearly independent vectors 
required to represent any vector in one-, two-, or three-dimensional vector 
space. 

We generalize this relationship by saying that in an r -dimensional 
linear vector space every vector x can be represented in the form 

x = ki&i + A 2&2 d~ ■ * * + (13-8) 

where a t> a 2 , a r , is any set of n linearly independent vectors. It 
follows from (13-8) that in such a space every set oi more than n vectors 
is linearly dependenl . 

We shah call a given set of n lineally independent sectors the base 
vectors (or the basts) of the //-dimensional linear \ eel or space, and we 
shall term the numbers ikuk 2 , . . .,k n ) the measure numbers associated with 
the basis a^, a 2 . • . • , a rl . 

In Sec. 3 we noted that every vector V in three-dimensional vectorspaeo 
can be represented uniquely by taking as a basis any set oi three linearly 
independent vectors a, b, c. But we saw that a special M*t of mutually 
orthogonal unit vectors i, j, k when used as a basis great h simplifies the 
calculations. This suggests the desirability of representing a vector x 
in the //-dimensional space in the form (13-8) m which the base vectors 
a, are the analogues of the unit vectors i, j, k The construction of an 
analogous set of base vectors requires the extension of the concepts of 
length and orthogonality to sots of vectors in //-dimensional space. In 
making these extensions we suppose that the scalar product a«b of a and 
b is a real number, and that a -a > 0 unless a - 0 Further the operation 
of scalar multiplication obeys the laws (1-3) and (1-3) 

We recall that in three-dimensional space two vectors a and b are 
orthogonal if a*b — 0 

and a is a unit vector if 

a*a - 1. 

We extend these definitions to vectors in //-dimensional space and show 
that when any sot of n linearly independent vectors ai, a 2 , . . a rt is given, 
one can construct a new set of vectors e*, e 2 , . . ., e n , such that 

e^e, = 0, if i j, 

= I, if i ** j . 


(13-9) 



320 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

A set of vectors satisfying the conditions (13-9) is called an orthonormal 
$et 

Let the set of vectors ax, a 2 , . a n be linearly independent, so that 
the equation 

+ <* 2&2 + • • • + a n a n = 0 (13-10) 

can be satisfied only by choosing = a 2 = * • * *= a n = 0. It follows 
from (13-10) that ai 0, for if it were a zero vector, the choice 

ot\ " 1 , a 2 » 0 , . . a n = 0 

would satisfy (13-10) and hence the vectors a* would be linearly dependent, 
thus contradicting our initial assumption. 

We shall write 

a**a t S5 |a t | 2 

and call |a»| the length of a,. Now denote the product of ai by the recip- 
rocal of its length |ai j by ei, so that 


ai 



Since ei*©i = 1, ei is a unit vector. The vectors 

e l i a 2> • • • > a n 

are obviously linearly independent. Consider next the vector 

e 2 = a 2 (a 2 *ei)ei. 

The scalar product e 2 *ei is 

e 2 * “ a 2*®i — ( a 2*®i) e i*®i * 0, 
since e* is a unit vector. Thus e 2 is orthogonal to ei and the vector 


t 

^2 



is a unit vector orthogonal to ©j. 

The set of vectors 

© 2 , a 3f . . a n 

is linearly independent, and we construct the vector 
e 3 — a 3 — (a 3 -ei)ei — (a 3 *e 2 )e 2 
which is orthogonal to both ej and e 2 . The vector 

t 

e 3 



is a unit vector, and the set of vectors 



SBC. 14] 


LINEAR VECTOR SPACES AND MATRICES 


321 


®1> ®4> • • •? 

is a linearly independent set. We continue the process by forming 

*4 *= ^4 — (a 4 *ei)ei — (a4*e 2 )e 2 — (a4*e 3 )e 3 

which is orthogonal to e u e 2 , and e 3 , and normalize it by dividing it by 
| e*| . The set of vectors 


®2> . • •> Rn 

is linearly independent, and a continuation of the procedure yields after 
n steps the desired set of orthonormal vectors 

©1> ®2) • • 'j 

14. Cartesian Reference Frames. When the base vectors i, j, k of Sec. 3 
are oriented along the xyz axes, the coordinates of their terminal points are 

i: (1,0,0), 

j : (0,1,0), 

k: (0,0,1). 

By analogy we can say that when a set of orthonormal base vectors ex, 
e 2 , . . ., e„ is oriented along “a cartesian reference frame in n~dimensional 
Euclidean space/' the terminal points of the base vectors have the co- 
ordinates 

e x : (1,0,0,..., 0), 
e 2 : (0,1,0, . . .,0), 

e 3 : (0,0,1,. . .,0), 

e n : (0,0,0, . . . ,1). 

In this reference frame every vector x has the representation 

x « xie t + x 2 e 2 H b z»e n , (14-1) 

where the x, are the components of x. 

On making use of the distributive law of scalar multiplication, we find 
that 

x-x = x 2 i + a-1 H h xl, (14-2) 

since e,-e ; = (14-3) 

where the symbol 6, y, the Kronecker delta, means 

Sn =1, if * - j, 

* 0, if is* j. 



322 ALGEfcRA AN!) GEOMETRY OF VECTORS. “MATRICES [CHAP. 4 

From (14-2) we conclude that the length |x| of the vector x is given by 
the formula 

I X I = + ‘*2 4" * * 4“ *n * 


This is the formula of Pythagoras in n-dnncnsional Euclidean space . 

Also, if 

y ~ !h e l + V' 2^2 + * • * f t/n^n, (14-4) 

then on forming the scalar product x-y \\c find 

x-y = X\\)i + x 2 y 2 4 h s n y n , (M-5) 

which has the same structure as formula (1-9). 

For the sum of two vectors x, y with components 

x: (.r lT .r L >,. . .,x n ), 

y: 0/l,//2,*. dJn\ 

we have the vector x + y with components 

x + y: (a + y u x 2 + y 2 , . . x n + ?/n), (14-0) 

and for the product of x by a scalar a, 

ax: (axi,ax 2 ,' • (14-7) 

If we have two vectors x and y in Euclidean three-dimensional space, we 
have a useful inequality 

(x-y) 2 < (x-x)(y-y) (14-8) 


which follows directly from the fact that 


cos 2 <9 


(x-y ) 2 

(x-x)(y-y) 


< 1 - 


We show next that the formula (11-8), known as the Cauchy-Schwarz 
inequality, is valid in an n-dimensional Euclidean space 
Indeed, 

(x-y ) 2 (x-y) 2 ] 

x*x — 2 4 — — 


(x-x)(y-y) - (x-y ) 2 = y-y 


y.y 


y-y 


— I y 1 2 



x-y 
y - 

y.y 


> 0 , 


which proves the inequality (14-8). We note that the equality sign in 
(14-8) holds if, and only if, y = 0 or x = ay for some scalar a. 

The formula (14-8) enables us to establish the result 


I x + y| < |x| + |y|, 


(14-9) 



SEC. 14] LINEAR VECTOR BP A CEB AND MATRICES 323 


analogous to the “triangle inequality” of Prob. 3 in Sec. 3. We compute 
|x 4 y| 2 = (x + yMx + y) = xx 4 y-y + 2x-y 



< |x| 2 4|y| 2 4 2|x-y|. 

(14-10) 

But from (14-8) 

(x-x)(y-y) > |x-y| 2 , 


so tii at 

Pl-|y|>|x-y|- 

(14-11) 


The substitution from (14-11) in (14-10) yields 

!x + yi 2 <!x| 2 +!y! 2 + 2 |x|.|y| = ( |x| 4 |y ! ) 2 , 
and on extracting the square root we get the inequality (14-9). 

In quantum mechanic and m several other branches of physics it is necessary to 
considei oi tiered sets of romplex numbers t ri,a 2, • . Such sets can he viewed as 
components of a vector x in an n-duneusionul complex vector spare . For the definition 
of addition of two complex vectors x, y with components 

x: f.ri,y 2 ,. • ■ ,?*), 

y- (Vl,//‘>, P/n), 

we can take formula (14-f>) and define the multiplication by a scalar a (real or complex) 
by (14-7). To make the length ,xj of tin* complex vector x real, we adopt as the defini- 
tion of scalar product of x and y the formula 


x * y * jp/i 4 ?2U2 4- • • 4 itt.Vn, ( » 4-12) 

in which J t denotes the conjugate of the complex number x t This foimula specializes 
to (14-5) when the components of vectors are real, since for real numbers i x = x t . We 
note from (14-12) that 

y-x * Xl C/i 4 1'iin ^ h JTnVn, 


so that x-y - y x, 

since the conjugate of the sum of complex numbers is equal to the sum of their conju- 
gates and the conjugate of the* pioduet is the product of the conjugates 
Formula (11-12) yields 

X X * XiX } 4 V2 H h JnXn, (14-13) 

so that I z I =* Vx’X is a real number 

The definition of linear independence of a set of complex vectors is that given in Sec. 
13 wheic the constants ot t are now in the field of complex numbers. 


PROBLEMS 

1. If one starts with the definition of a vector x as an n-tuple of n real or complex 
numbers (j*i,j- 2 ,- • .,2*n) and uses for the definition of sum and product the formulas 

x 4 y: Ui 4 ?/j, -..,^4 I in), 

kx: (kx i,. . .,Ax„), 

n 

x-y « 2 -V/*, 

t-1 



324 

then 


algebra and geometry of vectors, matrices [chap. 4 
(x + y)-z « x-z + y-z, 
x-(y + z) - x-y +X-X, 

(kx)-y - £(x-y), 
x-(ky) = k(x-y). 

S. Prove that if a (l) , a (!> , .... a (n) is a set of n linearly independent vectors in a 
complex n-dimensionai vector space, then the only vector x orthogonal to each of the 
vectors a <l> is the zero vector. 

3. Prove that a set of mutually orthogonal vectors is always linearly independent. 

4. Modify the proof of orthogonalization in Sec 13 so that it applies to a set of linearly 
independent complex vectors. 

15, Summation Convention. Cramer’s Rule. In dealing with expres- 
sions involving sums of quantities it is often useful to adopt the following 
summation convention: If in some expression a certain summation index 
occurs twice, we omit writing the summation symbol 2 and agree to sum the 
terms in the expression for all admissible values of the index. 

3 

Thus in a linear form X a t x x the summation index i appears twice under 

% «* i 

the summation symbol 2, and we shall write o,x t to mean OjXi -f a 2 x 2 

3 

+ a 3 x 3 . The symbol X “ a n + a 23 + <233 will be written simply 

t«* 1 

as a u . Again, a double sum 
3 3 

X £ Oq*»*y * <*11* 1*1 + <>12*1*2 + 013X1X3 + o 2 iX 2 x, -f 022*2X3 

+ 0 23 X 2 X 3 + 031X3X1 4* O32X3X2 + 033X3X3, 

which has two repeated summation indices i and j under the summation 
symbols, will be written as 

OjjXjXj. 

The range of admissible values of the indices, of course, has to be specified. 
Thus, the expression 

Ji- 1,2.3, 

a,jXh l j = 1, 2, 3, 4 

represents three linear forms 

OnXi + Oi 2 X 2 + O13X3 + 014X4, 

0 2 lX 1 + O22X2 + O23X3 + O24X4, 

<>31*1 + 032*2 + Oaa*3 + 034*4, 

corresponding to the three possible choices i » 1, 2, 3 of the free index i . 

, The summation index j is often called the dummy index because it can be 
replaced by any other letter having the same range of summation. The 



SBC. IS] UNEAR VECTOR SPACES AND MATRICES 325 

dummy index is analogous to the variable of integration in a definite 
integral, which can also be changed at will. Thus 


QijX-iXj — GkrXkXff 

it being understood that the indices i> j, r range over the same sets of val- 
ues. Unless a statement to the contrary is made, we shall suppose that 
the indices have the range of values from 1 to n. We shall thus write 
formulas (14-4) and (14-5), for example, as 

y = y^i, 


and (14-13) as 


x *y ** Wif 


x*x = £ t x t . 


We shall make use of this summation notation among other places in writing 
formulas for the product of determinants and for the expansion of deter- 
minants. 

We recall that a determinant 



a n 

012 ’ 

*01 71 

a %J | — 

<*21 

022‘ * 

‘02 n 


0nl 

0n2 * * 

* 0rm 


of order n represents an algebraic sum of n\ terms formed from the ele- 
ments a tJ in such a way that one, and only one, element from each row i 
and each column j appears in each term. 1 

The product of two determinants | a l} | and | b t j | , each of the order n, 
can be written as a single determinant |r t; | of order n in which the ele- 
ment c t} in the ith row and jth column is 2 

Cij 8=5 CL x kb)k ~ CLtlbji + 0i2^j2 "T * * * *4“ ^mbjn- (15-2) 

Inasmuch as the value of the determinant j 6* 7 1 is unchanged when its 
rows and columns are interchanged, the value of the determinant 

I C ij I 2=1 \ a i)\ | 6 .j| 

with the elements (15-2) is the same as that of the determinant |c*/| with 
the elements 


Cij ** ‘kbfcj &i\b\j T* 0>i2^2j "f" ' * * T* CLi n b n j. (15-3) 

1 A discussion of determinants is contained in Appendix A. 

* Since the number n k fixed, the term a tn b 3 „ does not represent the sum of terms with 
respect to n. Here n knot a summation index. Cf. Appendix A, Formula (1-10), 



326 


ALGEBRA AND GEOMETRY OF VEOTORvS. MATRICES [CHAP. 4 


If the oof actor 1 of the element a %3 in the determinant (15-1) is denoted 
by A ijy we can expand a, 3 in terms of the cof actors of elements in any 
row or column of the determinant. A reference to (1-5) in Appendix A 
will show that the following formulas include the Laplace developments 
of (15-1): 


(1%)* 1 ik 
a? 7-1 ki 


SjMtf 

&jk», 


(15- 


*-4) 
(15-5) 


where S 3 k is the Kronecker delta and a stands for the value of | a LJ j ; for 
if in (15-4) k 9^ the expression a tJ A,f- represents the sum of products 
of the elements in the jth column by tlie cofactors of (Ik* elements in the 
kth column. The value of such a sum is zero, since ll represents the 
expansion of a determinant with two like columns. If j = /;, the sum 
a tJ A ik is the sum of pioducts of the elements m thejth column, by the co- 
factors of those elements, vieldmg the value a — \a tJ 1 . Similar statements 
apply to (15-5) if we replace the word “column” by “row.” 

Formula (15-4) enables us to give a compact derivation of Cramer’s 
rule for solving a system ot n linear equations 


ci,jXj = b t (15-6) 

in n unknowns x, 

We multiply both members of (15-6) by the cofactors T,*. and sum with 
respect to i. We get 

ctijA t kX 3 “ A ikh t . 

But by (15-4) this is 


d jk axj = A lk b l . 


Tlie sum h 3 kX 3 = Xk, and we conclude that 

Ajklh 

xk ~ 

a 


(15-7) 


whenever a ^ 0. The numerator in (15-7) is the determinant obtained 
by replacing the elements in the Ath column of L/ j; | by the h t . The reader 
finding the foregoing calculations too concise will find a more expansive 
discussion in Sec. 2 of Appendix A. 


PROBLEMS 

1. Write out the following expressions in full. 

(a) 6 tJ a t ; (b) M a ub * (<0 (*) J dr lf (/) ^ dr/; (</) u tl ; (h) 

dx x 0.i a , 

ai « Z ~ bj; ( 1 ) a tJ a t k - S/k’, 0) (A) (() a l /r/, -= b x The symbols 5,, 

dXj 

denote the Kronecker deltas. 

1 Bee Appendix A. We recall that th»* eofactor of a t} is the signed minor M tJ of the 
element a, u the sign being ( — 1)' 



SEC. 16] LINEAR VECTOR SPACES AND MATRICES / 327 

2 . Write out the determinants represented by the expansion u®Aa and a$ % An, where 
A i, is the cofartor of the element a tJ in \a X3 1 . Also write out the determinants represented 
by a&A t 2 and n^A t ^. 

Z. Expend the determinants: 

Gil '02 0 13 Oi4 

0 fl22 «23 «24 

(<0 f\ A 

0 0 aw U34 

0 0 0 044 

111 a! 0 0 

(c) 1 1 x,\ ; (d) a>2 0 

.r'f sj jr£ 03 b-\ r { 

4. Multiply the determinant (h) in Prob 3 by the determinants (r) and (d). 

16. Matrices. In this section we introduce the concept of a matrix and 
discuss some rules of operation with matrices which are of value in the 
study of linear transformations. 

An m X n matrix is an ordered set of mn quantities a XJ arranged in a 
reel angular array of m rows and n columns. If m ~ «, the array is called 
a square matrix of order n. The quantities a l} are called the elements of 
the matrix. Thus, a matrix is an array 

On a 12 * ’ 
a :} a >2 * 02n 


tlmi a m2 ’ * * U m n 

where paren theses are used to enclose tin* array of elements. We shall 
denote matrices by capital letters, or when it is desired to exhibit a typical 
element of the matrix (16-1), we shall write {a a ). 

If the order of the elements in (lb- 1 ) is changed, or if any element is 
changed, a different matrix results For example, a triple of values (a l} 
] representing the cartesian coordinates of a point is a I X 3 matrix. 
If X a 2 , the matrix obviously represents a different point. 

Two m X n matrices A ~ (a u ) and B - (b tJ ) are said to be equal if, 
and only if, a tJ = b i} for each ? and j. That is, A - B only w T hen the 
elements in like positions of the two arrays are equal. 

We define the sum .1 + /> of two m X n matrices A = (<o ; ), B = (h l3 ) 
to be the array 

A+B = (« f . + ft*,), (H>-2) 

and their difference A — B to be the array 

A — B ~ {a tl — btj). 

We shall agree to say that the product of the matrix A ~ (a tJ ) by a con- 






328 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

giant k , written hA, is a matrix each of whose elements is multiplied by fc. 
Thus kA « (fca, ; ). 

If we have an m X n matrix A and an n X p matrix B ) we define the 
product of A and B t written AB , by the formula 

AB = (a ts b ik ), (16-3) 

where, as agreed in Sec. 15, the repeated index j is summed from 1 to n. 
Thus, the product AB is an m X p matrix, and we can multiply two 
matrices only if the number of columns m the first factor is equal to the 
number of rows in the second. 


Example 1. If 


and 


ab < 


p 5) 


a + 2 


and 
0-1 2 + 1 


/(1)<2) + (0) (0) + (2)(1) (I)(~ 1) + 


V0)(2) 4 (5) (0) 4- (6)(1) (OX- 1)4 



f 2 

-1 1 

R - 

0 

1 A 


Vi 

-2 -] 

\ 

/3 

-1 3X 

Oh 

2 

0 5 J 


\\ 

3 5/ 

b (2) ( - 

2) 

0)0) 4- 

1 (3)(- 

2) 

(2)0) 4- (- 

< 6 )(- 

2) 

(0)(1) 4- 




(oun t* (2)(— 1)\ 


(5) (2) -f (6)(-l)/ 



Also, if 


then 


AC 


C - 



/I 0 2w 2\ /(l)(2)4 (0)(3) +(2)(-lK / 0\ 

(2 -1 3 J f 3 } *= ( (2)(2) 4 ( — 1)(3) 4 (3)( — 1) ) ** ( —2 )• 

\0 5 6/ \ — 1 / \(0)(2)+ (5)(3) 4 (6)( — 1)/ \ 9/ 


We observe that the rule (16-2) for the addition of matrices requires 
that A 4* B = B 4 A } but it does not follow from (16-3) that the order 
of factors in the product AB can be interchanged even when the matrices 
are square. Indeed, for 



the rule (16-3) gives 

;). whii. (° 


Thus, the multiplication of matrices, in general, is not commutative. 



LINEAR VECTOR SPACES AND MATRICES 


329 


SEC. 16J 

However, if we have two square matrices of order n which have zero 
elements everywhere except possibly on the main diagonal, then it follows 
from (16-3) that 


fax 

0 


0 


0 ••• 

°\ 


0 • 

•' °\ 

ja 1 b l 

0 

•' °\ 

fig ’ * • 

0 


t> 2 - • 

-i 

0 

a^b 2 


o 

aj 

\o 

0 • 

.. J 

0 

0 • 

• • o, n bJ 


Such matrices are called diagonal . 

Thus for two diagonal matrices A and B y 

AB « BA, 


A diagonal matrix in which all elements along the main diagonal are equal 
is called a scalar matrix. A particular scalar matrix 


/i °--' 0 \ 
0 1 ••• 0 

\o 0 • • • 1 / 


(16-4) 


is called the identity ( or unit) matrix . 

We note that if I is the identity matrix and A is any square matrix, 
then 1 

I A - AI ~ A. (16-5) 

By analogy with the rules of ordinary algebra, we define the zero matrix 
0 to be the matrix such that 


O + A * A. 


It follows from (16-2) that all elements of the zero matrix are zeros. We 
observe that the product of two matrices may be a zero matrix even when 
neither of the factors is a zero matrix. Thus, if 


A « 


d 1 
0 0 
^0 1 


and 




then 


/0 0 (X 

AB = | 0 0 0 V 

V) o <y 


1 More generally we can show that if AX XA for every matrix A } then X is a scalar 
matrix. See Prob. 6. 



330 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

If the matrix is square, it is possible to form from the elements of the 
matrix a determinant whose elements have the same arrangement as those 
of the matrix. This determinant is called the determinant of the matrix. 
From any matrix, other matrices can be obtained by striking out a number 
of rows and columns. Certain of these matrices will be square matrices, 
and the determinants of these matrices are called determinants of the 
matrix. For an m X n matrix, there are square matrices of orders 1, 2, 
. . . , p, where p is equal to the smaller of the numbers m and n. 

Example 2. The 2X3 matrix 

4 m / a u a l2 tfl3\ 

\«21 «22 <*23/ 

contains the first-order square matrices (an), (a^), ( 023 ), etc., obtained by striking out 
any two columns and any one row. It also contains the second-order square matrices 

( «n &is\ /an flii\ /ai 2 <Ji3\ 

«21 022/ * \0n O23/ ' V022 o 23/ 1 

obtained by striking out any column of A. 

In many applications, it is useful to employ the notion of the rank of 
a matrix A. This is defined in terms of the determinants of A. A matrix 
A is said to be of rank r if there is at least one r-rowed determinant of A that 
is not zero, whereas all determinants of A of order higher than r are zero or 
nonexistent. 1 


Example 3. If 


/ 1 0 1 3\ 

A m ( 2 10 —2 J, 

V-l -1 1 5/ 


the third-order determinants are 


1 0 1 
2 1 0 
-1 -1 1 

1 1 3 

2 0 —2 

-1 1 5 


0, 

0, 



0 

1 

-1 



0, 


0 1 3 
1 0 —2 
-1 1 5 


0. 


Since 


1 0 
2 1 


3* 0, 


there is at least one second-order determinant different from zero, whereas all third-order 
determinants of A are zero. Therefore, the rank of A is 2. 

It should be observed that a matrix is said to have rank zero if all its 
elements are zero. 


1 Cf. Appendix A, Sec, 2. 



LINEAR VECTOR SPACES AND MATRICES 


331 


SEC. 16 ) 

If A * {dij) and B = (b tJ ) are two square matrices, then 

AB = (a ik b kj ) 

and the determinant of the matrix AB is 

\AB\ — | a x khkj | . ( 16 - 6 ) 

We note with reference to (15-3) that the elements in the zth row and jth 
column of the determinant in (16-6) are precisely those that appear in 
the product of two determinants | A | = | a tJ | and \B\ = \b x} \. Thus 

\AB\~\A\-\B\, ( 16 - 7 ) 

or in words, the determinant | A B | of the product of two matrices A and B 
ts equal tx> the product of determinants |.1 | and 

It follows from (16-7) that whenever the product of two matrices is a 
zero matrix, then the determinant of at least one of the factors is zero. 

A square matrix whose determinant is zero is called a singular matrix . 


PROBLEMS 


1. Make use of the definitions in Sec. 10 to establish the following theorems fot 
matrices: 


(a) A + B - B + A; (b) (A -f B) + C « A + (B + C); 

(c) (A + B)C - AC + BC ; (d) (\A + B) - CA + CB. 


2 . Verify that the matrices A and B in Example 1 of this section do not commute. 

3 . Multiply: 


(a) 


1 2 3 
3 1 2 
1 3 2 



(/>) 


1 2 3 

3 1 

1 3 2 


0 

0 

0 



4 . Show that (AB)C - A(BC). 

5. Determine the ranks of the matrices: 


1 2 3 

1 4 2 

2 6 5 


/l 0 1 

i? - { 0 0 1 

U 1 1 



D — 


-4 

1 



Is AB « BA? Is AE — EA? Are these matrices singular? 

6. If AX — XA for every matrix A, show that X is a scalar matrix. Hint: Let 
X — (x t j), then since AX - XA, a %) x t k - x^Uji for all choices of a tJ and a 3 Now choose 
«»/ * where $,(p> and b } ^ are the Kronecker deltas and p and q have fixed but 



532 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

arbitrary values ranging from 1 to n, and conclude that ■» xy, 0 if i j and 
a ? % i ** Xjj for each i and j, 

17. Linear Transformations. The matrix notation introduced in the 
preceding section enables us to study effectively properties of linear trans- 
formations. 

A set of n linear relations 

Vt ~ iy j == 1 ) 2 , . . . 7 fly ( 17 - 1 ) 

where the a», are constants, defines a linear transformation of the set of 
n variables x x into a new set y x . 

We can regard the quantities x iy x 2 , . . x n as components (or measure 
numbers) of some vector x referred to a set of base vectors a*, a 2 , . . a n 
in the n-dimensional vector space. The quantities y i} y 2 , . . y n can be 
viewed as components of another vector y referred to the same basis. 
The relations (17-1) then represent a transformation of the vector x into 
another vector y. Since the lengths of x and y and their orientations 
relative to the base vectors a, are different in general, we can look upon 
the transformation (17-1) as representing a deformation of space. 

When the components of x and y are represented by the column ma- 
trices 


n 


H 

* 

j ; 

Y «- 

f */2 1 

\j 

■ 

\,j 


the set of relations (17-1) can be written in the form 

Y * AX, (17-2) 

where A = (a XJ ) is the matrix of the coefficients in the linear transforma- 
tion (17-1) and the product AX is computed by the rule (16-3). 

If A is a nonsingular matrix, we can solve Eqs. (17-1) for the x % by 
Cramer’s rule (15-7) and obtain the inverse transformation 





(17-3) 


where A x j is the cofactor of the element a l} in the determinant a » |a, 7 | 
erf tjie matrix A . 

The set of equations (17-3) can be written in matrix notation as 

X « A~ l Y , 



sec, 17] 
where A~ l is 


UNEAR VECTOR SPACES AND MATRICES 


333 


1 ^ 
1— * 

I *-* 

A21 

A n \\ 


a 

a 

a \ 


f An 

1 ^ 

I W 

^4n2 

S3S 

a 

a 

a 


\ Ain 

A2n 

Ann j 

' a 

a 

a 


(17-4) 


It is natural to call A~~ x the inverse matrix of A. We note that the 
inverse matrix can be constructed whenever A is nonsingular, that is, 
whenever the determinant \A | m a does not vanish. 

If we form the product of A and A “” 1 

AA~ l = (a, k (17-5) 

and recall 1 that 

atkAjk = 5 X jC t, 

we can write (17-5) as 

A A~' = («„) = I, (17-6) 

where I is the identity matrix. 

Since the determinant of the product of two matrices is equal to the 
product of their determinants, we conclude from (17-6) that 

\A- i A\ = \A~ l \-\A\ = \I\= 1, 


so that 


1 

Ui 


(17-7) 


Multiplying (17-6) on the left by *4“ l and on the right by (A~ l )~ l gives 

A~ X A = 1 ~ AA~ l . (17-8) 

In addition to the inverse matrix A “* 1 we shall make frequent use of the 
matrix 

l<l\l #21 * * * #nl\ 

#12 #22 ' * • a n2 


(17-9) 


V* 


In #2 n ' ' ' ei n 


1 See (15-5), but note the relation of the subscripts on the to the rows and columns 
in (17-4), 



334 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

obtained by interchanging the rows and columns in the matrix 


j a ll O] 2 * * * #uA 

#2! #22 * * * #2n 

'#/il a n‘2 * * * find 


(17-10) 


The matrix A' is called the transpose of A, 

On using the laws of addition and multiplication of matrices it is easy 
to show that 

(A + By « A' + B\ 

(k AY - AvT, 

(Any - b’a\ (i7-ii) 

(Note order.) 

If we recall the relation (17-8), 

A~ l A - AA~\ 

and form the transpose 

(A~ l A)' = (AA~ l y, 

w© get, on making use of (17-11), 

A'{A- l y - (A~ l yA f . (17-12) 

Multiplying both members of (17-12) on the left by C4')~\ we get 

(d'l^d'Ol- 1 )' = (ri'r^i” 1 )'.!'. 

Hence (A~ l y = (A')~ l (AA~ 1 )' « Cl')" 1 . 

Thus (A~ l ) f = WO” 1 . (17-33) 

The important result embodied in (17-13) slates that the inverse of the 
transpose of the matrix A 'is equal to flu Inuisposi of its inn esc. 

In many calculations it is necessary to compute the inverse of the 
product of two nonsingular matrices A and B. We can obtain the desired 
result as follows: Since 

(AB)(AB)~ X = /, 

or (see Prob. 4, Sec. 10) 

AB(Ali )- 1 = I, 

wo get, on multiplying both members of this relation on the left by A -1 , 
A~ l AB{AB)~ l = /l"" 1 
or B(AIS)~ X = A~\ 

Multiplying this result on the left by B~ l , we get the desired result 

{AB)~ l = B~ X A~ X . 


(17-14) 



SEC. 17] lil NEAR VECTOR SPACES AND MATRICES 335 

(Note order.) This result can be extended in an obvious way to more than 
two matrices, so that, for example, 

(ABC)” 1 ~ 

Example 1. Compute A~ l for the matrix 


-a -> 


A n m ~1, A 12 * —3; An * —2, A 22 ~ 1. 


Since a *» I A 1 


We note tlmt | A 1 |= s ~'H®'1/bt|. 

Example 2 If A is a iionsingulai matrix, show that the matric equations 

AX • I and XA « 1 

have unique solutions A T = A ~~ l . 

On multiplying both members of the given equations by A” 1 , we get 
A- 1 AX « A" 1 / and XAA~ l « IA~K 

But A -, A ~ A A “ l ~ / and A" 1 / - /A"" 1 - A~\ 


If we have two successive linear transformations 


Vi = flub, 


(17-15) 


the direct transformation from the variables .r t to the z, is obtained by 
inserting for the y } in the second set of Eqs. (17-15) from the first set. We 
thus get 

2* = bt/ijkT a. (17-16) 

The transformation (17-16) is called the product of the transformations in 
(17-15). If the variables (jr^r^, . . . , r M ), O/i AJh • ■ • djn), and (z u z 2 ,. . .,x„) 
are interpreted as components oi the vectors x, y, z, represented by column 
matrices 





336 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

we can write Eqs. (17-15) as 


Y m AX , 
Z ^ BY, 

and the product transformation (17-16) as 

Z = BdX. 


(17-17) 


(17-18) 


Thus , when the variables are subjected to a linear transformation (17-15) 
with a matrix A and the variables are subjected to a linear transformation 
with a matrix B, the product transformation has the matrix BA . Since BA 
in general is not equal to AB , the order in which the transformations are 
performed is material. 

When it is desired to interpret Eqs. (37-15) as transformations on the 
components of the vectors x, y, z, Eqs. (17-17) and (17-18) can be WTitten 
in the forms 

y = Ax, 


z - By, 


(17-19) 


z - BAx } 

where x, y, z are regarded as the column matrices X , F, Z, respectively. 
The matrices in Eqs. (17-19) can be viewed as operators transforming a 
given vector into another vector. Since 

A (kx) = kA x, k const, 

and A(x + y) « Ax + Ay, 

one often speaks of A as a linear operator. 


PROBLEMS 



find A~ l and Verify that (AB)' - B'A' and (AB)~~ l * J 5 ~ 1 A~ 1 . 

2 . Prove that (ABC)' - C'B'A'. 

3 . Prove that (A “ 1 )~" 1 ** A. 

4 . Prove that if A is singular, there exists no matrix B such that AB » /. 

5 . If yi » zj cos a — X2 sin a, 1/2 »* £1 sin a + X2 cos a, find A~ l t A' and show that 
A "" 1 ** A*. If x is a vector with components (r^a*), what is the geometric relation of 
i toy? Write out the inverse transformation x « A ” 1 y. 

6 . If yx ** x% - £2, 1/2 « *f *2, what is A*" 1 ? Is A ” 1 « A'? If x is a vector with 
components (xi,X2)> what is the geometric relation of x to y? 



SBC. 18] 
7. If 


LINEAR VECTOR SPACES AND MATRICES 


337 


i , i 

Vl “^ Xl + ^ Xi ’ 

yt «• 

compute the matrix A ' for the inverse transformation and compare it with the given 
matrix A. If x is a vector with components ( x\,x%,xz ) and y is a vector with components 
(Vi.I/ 2 , 1 / 3 ), what is the geometric relation of x to y? 

8. Let 


and consider the vector 

a -G a) 


y » Ax. 

Compute x « A* x j. Is it true that A* A l ? 

9. If 

yi = T\ cos a -f xi sin a, 


y 2 * ~~x\ sin a + cos a, 

and 

Zi ** j/i cos 0 + 2/2 sin 0, 


» — 2/1 sin 0 -f y 2 cos 0, 

find the product transformation directly and also by computing the product of the 
matrices as in (17-18). Compute BA and AB, A~ l B~ l , and (BA)* 1 . Also find (BA)' 
and compare it with (BA)* 1 . 

10. If 

yi « 2 ji -f X2 , 

1/2 * *1 ~ * 2 , 

and z\ ** y\ - 2 / 2 , 

Z2 * 2 y\ + 2/2, 

perform the calculations required in Prob. 9. 

18. Transformation of Base Vectors. In the preceding section we in- 
terpreted the set of linear relations 

Vi ~ Gift] (18-1) 

as transformations of components (xift 2) . . x n ) of a vector x into com- 
ponents (yi,t/ 2 r * * *,2/n) of another vector y when the vectors are referred 
to the same basis (a,*), so that 

x * x*a, and y ® y t &i. (18-2) 

If we introduce a new system of base vectors o tJ obtained from the set 
a, by a linear transformation 



338 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

the vectors x and y in the new reference system will have certain representa- 
tions „ /to 4 K 

x = £*«*, y * v i«i* (18-4) 

We raise two questions: (1) What is the relation of the components of 
vectors in the two representations (IS-2) and (18-1) when the base vectors 
are transformed by (18-3)? (2) What is the form of the transformation oi 
the components £, into *?, which corresponds to the deformation of the 
vector x characterized by Eqs. (18-1)? 

To answer the first question we insert from (18-3) in (18-1) and get 

x - Ihj^sLj, (18-5) 

while a reference to (18-2) shows that 

x = a,.r, “ Si r rj. (18-6) 

From (18-5) and (18-0) we cone) ude that 

•r, ~ M*. U# 7' 

This formula is the desired relationship connecting the < <>mponcnU of x 
when it is referred to two different base systems i da ted bv r \ 18-3 

We note that in the transformation (l<S-3) the summation is on tlie 
second index.; while in (18-7) it is on the first index. In other words, the 
matrix of coefficients b tJ in (18-7) is the transpose of the matrix (b l} ) in 
(18-3). 

If we write the matrix in (18-3) as 

(b tJ ) - H, 

the set of equations (18-7) e in be wi it ten as 

x - />"|, (18-8) 

x and | being the column matrices with components (t’i r r 2} . . .,r w j and 
(£bi;2> • • • i £?»)■ 

On multiplying (18-8) by (li , )~~ l on the left we get the solution for £ 
in the form 

$=(«';- ‘x. 

Formulas (18-8) and (18-9) give a complete answer to the first question 
The relationship connecting the components (;/i,;/ 2 , . . , } y n ) with (?? l3 
i) 2 j * . -jVn) can be represented similarly by 

y = B't\ and -q - (ft'rV (18-10) 

We proceed next to the answer of the question concerning the form of 
the deformation of space (18-1) in the new reference frame a t . 

We write Eqs. (18-1) in matrix form as 

y - Ax, 


(18-11) 



SBC. 18] UNBAR VECTOR SPACES AND MATRICES * 339 


substitute for x from (18-8) and for y from (18-10), and obtain 

B' 1] - AB'b (18-12) 

To solve for r\ we multiply on the left by ( B')~~ l and get 

tl = (18-13) 

Thus the relationship between the components (fi,$ 2 ,...,£«) and (r}i,V 2 > 
. . ,,r) n ) is determined by the matrix 

8 m (B')~ l AB'. (18-14) 

Since the matrix S characterizes the same deformation of space as the 
matrix A , the matrices A and S related in the manner of (18-1 1) are termed 
dmilar . To avoid carrying primes, we set B' ~ (\ and formula (18-14) 
(hen assumes the form 

S m C~ l AC, (18-15) 

and (18-13) becomes 

t, - SI (18-10) 


One of the important problems in the theory of linear transformations is 
to determine a reference frame in which the equations tor the deformation 
of spare assume forms v hub admit of simple interprets ions. For example, 
if it proves possible to find a matrix C such that the matrix S in (18-15) 
has the diagonal form 


S =- 


then Eq. (18-10) shows that 


/X, 

o ■ 

°\ 

0 

x 2 ••• 

.* 



• \J 

Vl 

->J7 

il 


V2 



Vn 




(18-17) 


In three-dimensional space these correspond to simple elongations (or 
contractions) of the components of the vector in the directions of base 
vectors a t determined by the matrix C ~ B f [see (18-3)]. 

Whether or not a matrix C reducing A to the diagonal form S can be 
found clearly depends on the nature of deformation specified by A. In 
many problems in dynamics and in the theory of elasticity, the deforma- 
tion matrix A will turn out to be symmetric, and we shall see in Sec. 20 
that such matrices can always be diagonalized by finding a suitable ma- 



340 AliGEBRA AND GEOMETRY OP VECTORS. MATRICES (CHAP. 4 

irix C. This fact turns out to be of cardinal importance because it enor- 
mously simplifies the analysis of many problems. 

In the following section we shall study properties of the matrix A in 
(18-1) for those transformations that leave the length of every vector x 
unchanged. In three dimensions such transformations represent rotations 
and reflections. 

19. Orthogonal Transformations. Let us refer our ?i-dimensional space 
to a set of orthonormal base vectors e2> . . ©n, introduced in Sec. 13. 
Relative to this basis the vector x has the representation 


x = e&i 


and its length |x| can be computed from the formula 

| x | 2 = x t x t . (19-1) 


Let us investigate the structure of the matrix A in the class of transforma- 
tions / 1 f \ r>\ 

Vi * o tJ x } (19-2) 

which leave the length |x| of the vector unchanged. Now, the square 
of the length of the vector y is 

lyi 2 = 09-3) 

and since we suppose that | x | « | y | , 

y<Vi « x % Xi. (19-4) 

We insert in (19-4) from (19-2) and get 

(a %j Xj)(a ik Xk) * x,x t 

or <itja t }rXjX k *» d^x *, (19-5) 

since &i&jXi k = *kk = x t x { . 

On equating the coefficients of Xjit in (19-5), we get the set of restrictive 

conditions t /ir4 

d% j&lk §)k (I9~(i) 


on the coefficients a t; if the transformation (19-2) is to leave the length of 
every vector unchanged. 

Equations (19-6), when written out for n = 3, are 

a ii 4" a ti 4* &ii ** 1, 

a 12 + d %2 + U32 ~ 1, 
a l3 + &2Z 4* O 33 335 1> 

Ol2 a l3 4* ^22^23 4“ U32O33 ** 0, 

013^11 4* «23&21 4* G33U31 = 0, 
a u ai2 4“ G21G22 4- <*31^32 m 0. 



341 


SBC. 19 ] LINEAR VECTOR 8FACES AND MATRICES 

The determinant of the matrix in (19-6) is 

I diflik I = | Sjk | = 1, (19-7) 

and if we recall the rule for multiplication of determinants [cf. (15-2)], 
we conclude from (19-7) that 

I aifiik I = I a,j I • I aij I = a 2 - 1, (19-8) 

where a is the determinant of (ay). Equation (19-8) states that 

d = dr 1 . 

In three dimensions the situation when a = 1 corresponds to a rotation 
of space relative to a set of fixed xyz axes determined by the unit vectors 
i, j, k. The circumstance when a « — 1 corresponds to a transformation 
of reflection (say, x = -~x, y « ~~y, z ~ — z ) or to a reflection followed by 
a rotation* 

A transformation (19-2) in which the coefficients o t j satisfy (19-6) is 
(‘ailed an orthogonal transformation; it is called the transformation of rotation 
if |a i; j = 1, whatever be the dimensionality of space. 

If we denote by A ' the transpose of ( a tJ ) ~ A in (19-2), we can write 
the orthogonality condition (19-0) in matrix form as 

A' A « I. (19-9) 

On multiplying this by A on the right we get 

A' ~ A - 1 . (19-10) 

Thus, in an orthogonal transformation the inverse matrix A~ l is equal to 
the transpose A ' of A . 

When Eqs. (19-2) are written in the form 

y - Ax, 

we can write their solutions for the x x as 

x « A~ l y. (19-11) 

We conclude from (19-10) that the solutions of Eqs. (19-2), when the 
transformation is orthogonal, are 


x % = ajiyj. (19-12) 

In Sec. 17, we saw that the matrix of the product of two linear trans- 
formations is the product of the matrices of the component transformations. 
Using this fact and the property (19-10) it is easy to show that the product 
of two orthogonal transformations is an orthogonal transformation . 



342 


ALGEBRA AND GEOMETRY OP VECTOkS. MATRICES 


[CHAP* 4 


PROBLEMS 


1. Verify that the transformations 


(a) 

1/1 

* xi cos a x* sin a, 


1/2 

* X\ sin a -}- X2 COS a, 

and 



(b) 


1 1 

2/1 

= VT 1 + ^ X3 ' 


V2 

= X-2, 



1 l 


.V3 



are orthogonal. Do they represent rotations? 

2. Discuss the transformation 




Vl ~ 3 V 2 il+ 3 h' l+ 


J/3 


2 2 1 
3" J+ 3" 2 “3 JS 


Find the inverse transformation, 

3 . Prove that the product of any number of orthogonal transformations is an orthog- 
onal transformation. 

4. Tf A i.N a symmetric matrix (so that A ' — A) and 8 is an orthogonal trariofoi mation, 
prove that the matrix B » S ~ is symmetric Thus, orthogonal transformations do 
not destroy the symmetry of A. 

6. J*‘t 

( 1 1 0 
1 2-1 
o-i a 

and let C be an orthogonal matrix 


f c it 

02 

c 1 A 

^21 

C22 

C23 1 


C32 

C33/ 


Write out the set of equations which the c„ must satisfy if C~ l AC * S, where $ is a di- 
agonal matrix 

6, Is the transformation 

Vi » 3xi — 3% 

3/2 ~2xi + 

orthogonal? Find the inverse transformation. Determine the components of x: (x h i 2 ) 
and y: (yi,yt) when the base vectors ex, e 2 are rotated through 45 and 90°. 



LINEAR VECTOR SPACES AND MATRICES 


343 


sec. 20] 

7. If y% «* OijXj is a linear transformation for the components of a complex vector 
x: (a?i,X2, . . . t z n ), which preserves the length |x| of the vector, show that dtjaik ** &jk 
or X'A ** I, where X is the conjugate matrix formed by replacing every element a,y of 
A by a X) . Transformations such that X' «* A ~ 1 are called unitary; they are of great 
importance in quantum mechanics. 

20. The Diagonalization of Matrices. We saw in See. 18 that the 
determination of a nonsingular matrix C such that the given matrix A 
reduces to the diagonal form S by a similitude transformation C~~ l AC is 
equivalent to determining a set of base vectors relative to which the trans- 
formation 

Vi = a t jXj (20-1) 

assumes the form 

Vl *= ^lft, V2 ** ^2$2j •••» Vn = (20-2) 

We thus seek a solution of the matric equation 

C~~ l AC - S (20-3) 

in which A = ( a XJ ) is a given matrix, C the unknown matrix 


Mi 

<12 * * 

* Mfc 

. . . 

C ln\ 


c=h 

c 22 * • 

* ^2 k 

• # * 

C2n 

j 

(20-4) 

V n l 

Cn2 ’ • 

' Cnh 


C nn l 


and S is the diagonal matrix, 






A, 

0 


°\ 



4 

X 2 

. . . 

1 

0 

• 

(20-5) 

\o 

0 


kJ 



On multiplying (20-3) on the left by C we get an equivalent matric equation 


AC = C*S, (20-6) 

provided that the solution of (20-6) yields a nonsingular matrix C. 

Now the matric equation (20-6) is equivalent to a system of linear equa- 
tions 

a t jCjk = Cikhk, no sum on k, k — 1, . . n (20-7) 

obtained by equating the corresponding elements in the products AC and 
CS. 

For every fixed value of h, the system (20-7) represents a set of n linear 
homogeneous equations for the unknowns (cu,c 2 fc,. . . ,c„*) appearing in the 



344 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

kth column of (20-4). The fact that the system (20-7) is homogeneous 
can be made plainer by rewriting it in the form 

(Uij — = 0, no sum on k. (20-8) 

We recall 1 that a system of homogeneous equations has solutions other 
than the obvious solution eq*, ~ C 2 k ~ ~ c n * — 0 if, and only if, its 

determinant 2 

|0 fJ - X6„| - 0. (20-9) 

On writing out this determinant in full, 

a n “ X Ul2 * ‘ * a ln 
<121 ^22 — X • • * &2 n 

* 0 , ( 20 - 10 ) 

Rnl Un2 * 4 ' Qnn X 

we see that (20-10) is an algebraic ecjuation of degree n in A. Accordingly, 
there are n roots of this equation, say X — Xi, X = X 2 , ...» X = X n , and 
corresponding to each root X = X^ (A = 1, the system (20-8; will 

have a solution 

(pik)<'2ki * ♦ • fnk)* (20- 1 1 ) 

The solution (20-11) yields the Ath column of the matrix C. If the roots 
\ x , X 2 , . . ., X„ are all distinct, one can prow* that the matrix C will be non- 
singular. 3 When the roots A* are not distinct, it is impossible, in general, 
to reduce A by the similitude transformation (20-3) to the diagonal form, 
because the desired nonsingular matrix C may not exist. In important 
special cases, however (for example, when A is a real and symmetric ma- 
trix), one can construct C such that has the diagonal form even when 
some, or even all, roots are equal. 

A brief discussion of this is contained in the following section. 

As a matter of terminology, Eq (20-9) is called the characteristic equation 
and its solutions are characteristic values of the matrix (a tJ ). The solutions 
(20-11) of the system (20-8) corresponding to these characteristic values 
are called characteristic vectors , 4 

1 Appendix A, Sec. 2, 

* Note that this determinants! equation when written in matrix form is | A — \f | ® 0. 

* Or in the language of vectors, if we regard each column of C as a vector c {k) : 
(ciktC 2 k f > • *,Cnk), the vectors c {k) (k « 1,2, . . n) will be linearly independent. A simple 
proof of tiiis is given in I. S. Sokolnikoff, “Tensor Analysis," pp. 33-34, John Wiley 

Sons, Inc., New r York, 1951. 

* The hybrid terms eigenvalues for the A* and eigenvectors for the c (&) : . ,,CnJt) 

Are used by some writers who do not mind mixing German with English. 



34S 


SEC. 20J UNEAR VECTOR SPACES AND MATRICES 

Example 1. Reduce the matrix 

A m ( oij ) * j) 

to the diagonal form S by the similitude transformation C~ l AC. 
The characteristic equation (20-9) hero is 


( 20 - 12 ) 


Its solutions are Xi *» 0, X 2 » 2. The desired matrix C in our case has the form 



the columns in which satisfy the system of equations (20-8), yielding 
(an — XjOci* 4 aiar^ * 0, no sum on k, 

021 O* 4 (022 — Xjt)c 2 i = 0 , k *» 1 , 2 . 

Since an — 1, 012 — — 1, <*2 1 ** — 1, 022 ** 1, we get, on setting k 1 and Xi 

c n - C21 0, 

~Qi 4 <’21 * 0. 


(20-13) 


(20-14) 

0, 

(20-15) 


As is always the case with nontrivial homogeneous systems of equations, 1 there fire 
infinitely many solutions of the system (20-15). If we set rn = a (any constant), Eqs. 
(20-15) give rut =* a. 

Thus the vector c m . ( 01 ,^ 21 ) appearing in the first column of (20-13) has the compo- 
nents oj — C 2 i ~ a. Since any matrix C accomplishing the reduction will do, we can 
take 2 a *■ 1. 

The substitution of k * 2 and Xj « 2 in (20-1 1) yields the system 

(1 - 2)fj2 - t’22 * 0, 

— C12 4(1 — 2)^22 ** 0, 


or 


— c 12 *— C 22 30 0. 


Again there are infinitely many solutions, and if we take c& * a, then c 2 2 ** —a. We 
can set a » 1 if we wish, so that the elements of the second column in (20-13) are C 12 *» 1^ 
022 « — 1 . The desired matrix C, thei efore, is 


C - 


(! 



The inverse of C is easily found to be 

C~ l *® 

so that C~ l A C is 

(H ¥t \ ( 1 -i\ /i 

\H -W ' V-i 1 / ' M 


Vi \ 

W 2 - y 2 ) ’ 



1 See Appendix A, Sec. 2. 

* Usually one normalises solutions so that the length ef the column vector c (fc) is 1, 
This would correspond to the choice of a *» l/\/2. since c\i 4 c\% - 1. 



346 ALGEBRA AND GEOMETRY OF VECTORS, MATRICES [CHAP, 4 

On multiplying these matrices we get 



as we should, since 



(20-16) 


as we knew from the start [see (20-5) J. 

If we interpret A as a matrix operator characterizing the deformation of a vector x 
into a vector y [see (18-11)], the result (20-16) states that in a suitable reference frame 
the components of x and y are related by 

vi =* 0*$i, m = 2&. 

We thus have a deformation of space corrasponding to the twofold elongation in the 
direction of one of the base vectors. In the notation of Sec. 18, C * so that one can 
actually write out Eqs. (18-3) for the transformation of the base vectors. This, how- 
ever, is seldom required because the essential matter is to determine the deformation 
characterized by A rather than a reference frame giving a simple form of the deformation. 

Example 2. Determine the characteristic values of the matrix 

/ 1 - 1 - ] \ 

(o„) - ^-1 1 -1 j- (20-17) 

The characteristic equation this time is 


1 - X 


-1 

-1 




(1 - X ) 2 - 3(1 - X) - 2 - 0. 


We easily check that the solutions of this cubic are X] = 2, Xg = 2, X 3 =*» — 1 . Since we 
have a double root Xi *» X 2 — 2, the solution of the system (20-8) will enable us to de- 
termine only two linearly independent columns of the matrix C. The matrix (20-17), 
however, is real and symmetric , and one can, in fact, construct the third column of the 
matrix C such that 

C~ l AC - 8, 

( Xi 0 0\ /2 0 0\ 

0 X 2 0 J * f 0 2 0 J • 

0 0 Xj/ Vo 0 - 1 / 

However, the theory presented in this section does not explain how this can be accom- 
plished. 


1. Diagonalize the matrix 


PROBLEMS 



and determine, in the manner of Example 1, the matrix C. Discuss the meaning of A 
when viewed as an operator characterizing a deformation of space. 



347 


SEC. 21] LINEAR VECTOR SPACES AND MATRICES 

2 . Find the roots of the characteristic equation for the matrix 



3. Find a matrix C reducing 

( 

4-{o- 

\ 0 

to the diagonal form by the transformation C 

4 . Diagonalize the matrix 

2 4 

4 2 

—6 —6 

6. Prove that the roots of characteristic equations of all similar matrices are equal. 
Hint: Write the characteristic equation of [of. (20-9)] in the form | C~ l AC — X/| 

- 0. But | C-'AC - X/ 1 - |C- l (A ~ X7)C| » |4 ~ A/ 1, since 1C" 1 ! « 1/|C|. 

21. Real Symmetric Matrices and Quadratic Forms. Let the matrix 
A = (a tJ ) in a linear transformation 

Vi = at]*,, hi ** 1, . . n, (21-1) 

be real and symmetric, so that A' = A (or a tJ * = a ;t )- We shall indicate 
that in this case the matrix A can always be reduced by the transformation 
C~~ l AC to the diagonal form S. Moreover, C can be chosen as an orthogonal 
matrix; that is, a matrix such that C"" 1 ~ C' [cf. Eq. (19-10)]. 

Linear transformations with real symmetric matrices dominate the 
study of deformations of elastic media. Real symmetric matrices also oc- 
cur in the study of quadratic forms 

Q(x h z 2 , . . . ,x n ) ss a X3 x % x^ h i * 1 , 2, . . . , n, (21-2) 

which arise in many problems concerned with vibrations of dynamical 
systems. 

We can always suppose that the coefficients in a quadratic form (21-2) 
are symmetric because every quadratic form Q can be symmetrized by 
writing it as 

Q = MK + 

= bijXtXj, 

in which the coefficients 

M ( a ij + a n) 

are obviously symmetric. Henceforth we shall suppose that our quadratic 
forms have been symmetrized so that a %J — 

It will follow from discussion in this section that the problems of reduc- 
tion of the transformation (21-1) with symmetric coefficients to the form 

= Xifi, 1)2 ~ . . * . 8=5 *nf« (21-3) 




848 ALGEBRA AND GEOMETRY OF VECTORS. MATRICES (CHAP. 4 

and of the quadratic form (21-2) to the form 

Q *= Ai£i + + • * * + X n fn (21-4) 

are mathematically identical. 

We first note several properties of quadratic forms. If the variables x% 
in (21-2) are subjected to a linear transformation 

x, « c ty (21-5) 

the form (21-2) becomes 

Q = <**,(<>**£*) fork) 

= aijf'ikf'jrt/cZr- 

We denote the coefficients of r by 5jk r , so that 

0 - bkrter, (21-0) 

where V * a l} c lk c ]r . (21-7) 

Since i and j are the summation indices and o tJ = a ;t , we see that the value 
of bkr is not changed by an interchange of k and r. Titus, \u» conclude 
that the symmetry of the coefficients in a quadratic form (21-2) is not 
destroyed when the variables x % are changed by a linear transformation 
(21-5). 

If we write (21-7) in the form 

bkr ~ CikittijCjr) > 

we see that the sum a XJ c Jr is an element in the zth row and the rth column 
of the matrix 

AC -/), 

or (dtjCjr) » (d,r). 

The product c % k{a l} c )f ) = c tk d ir is the element in the kth row and the rth 
column of the matrix CD. Thus we can write (21-7) as 

B - CAC . (21-8) 

The result (21-8) can be stated as a theorem. 

Theorem. When the variables x t in a quadratic form (21-2) with a matrix 
A are subjected to a hncar transformation (21-5) with a matrix C } the resulting 
quadratic form has the matrix CAC. 

If the linear transformation (21-5) is orthogonal, then C * C~~ l , and 
hence (21-8) can be written as 

» B - C~ l AC. (21.9) 

We conclude from (21-9) that the reduction of a symmetric matrix to the 



349 


SEC. 21 ] LINEAR VECTOR SPACES AND MATRICES 

diagonal form by an orthogonal transformation calls for a solution of the 
matric equation 

S - C~~ l AC. ( 21 - 10 ) 

This equation is identical with that considered in the preceding section. 
When the roots of the characteristic equation 

|«0 1 = 0 ( 21 - 11 ) 

are distinct and real, the method of Sec. 20 enables us to compute a matrix 
C which can be shown to be orthogonal. As a matter of fact, the desired 
matrix C can always be found whenever the matrix A is real and symmetric. 
Moreover, it can be shown that the roots of symmetric real matrices are 
invariably real. 1 

The fact that the columns of C are linearly independent can be established easily when 
A is symmetric and the X* an; all unequal. Let c be a characteristic vector for X and 
let c' be a characteristic vector for a different value, X'. Then, from (20-7), 

a t jCj «* Xr* and a tJ cj * XVj. 

Multiplying these equations by cj and c t , respectively, gives 

a ij c 'i c j “ Xclc, and Oj/y\ ~ XVjcJ 

after summing on i. Since a tJ ■* a ;i , the left sides of these equations are equal. Hence 
by subtraction, 

0 * \CiC t — X'c t ri » (X — X0c»cJ. 

Since X^X' we get c-c' *» « 0, so that the vectors c and c' are orthogonal and thus 

linearly independent. 

If the roots X, are all positive, Eq. (21-4) shows that the quadratic form 
(21-2) assumes positive values for all nonzero values of the variables 
Such quadratic forms are called positive definite, . They appear in numerous 
investigations in mathematical physics. 

An analogue of a symmetric quadratic form (21-2) in which the variables x, are com- 
plex is a bilinear form 8 

H «* OtfftXj ( 21 - 12 ) 

in which o,*y = & }% . Such forms are called Hermitian, and their matrices ( Oi } ) * A are 
Hermit tan matrices. Since a,/ *» dj %t it follows that the elements on the main diagonal 
of A are necessarily real and that 

A' - I. 

From the structure of (21-12) it follows that the Hermitian forms assume only real 
values for arbitrary complex values for on taking the conjugate of (21-12), we get 

H «* dijX x £j *= aji2 } Xi ** //, 

which proves that H is real. 

1 For proofs utilizing the notation of this section, see I. S. Sokolnikoff, “Tensor Analy- 
sis/’ pp. 37-40, John Wiley k Sons, Inc., New York, 1951. 

* Cf. Prob. 7, Sec. 19. 



350 


ALGEBRA AND GEOMETRY OF VECTORS. MATRICES [CHAP. 4 

Hermitian forms occur in quantum mechanics, and a discussion of the reduction of a 
quadratic form to a sum of squares (21-4) can be generalized to show that (21-12) can 
be reduced to the form 

H ®» Xi£i£j *+■ +■*** + Xft| n £ n 

by a linear transformation (21-5) with a unitary matrix C defined in Prob. 7 of Sec. 19. 

22' Solution of Systems of Linear Equations. In Sec. 15 we derived 
Cramer’s rule for solving the system of equations 

(lijXj = b>. ( 22 - 1 ) 

When the number of equations in (22-1) is large, Cramer’s rule is inefficient, 
since it requires evaluating determinants of high orders. For this reason 
all practical methods of solving the system (22-1) depend on reducing it 
by some process to an equivalent system whose matrix is sufficiently 
simple to enable one to compute the unknowns without great effort. 

The system (22-1) can be written in matrix notation as 

Ax = b, (22-2) 

where A * (a#), x is the column matrix (ji,t 2> . . . } x n ), and b is the column 
matrix (61,63, . • *,6 n ). If A is nonsingular, the solution of (22-2) is 

x = A~ l b (22-3) 

so that the determination of unknowns hinges on constructing the inverse 
matrix A "~ 1 . The development of effective methods for inverting matrices is 
a major problem of numerical analysis. One of such methods depends on 
a reduction of the system (22-2) to an equivalent system 

Bx - c, (22-4) 

in which B has the triangular form 



612 

1>1 3 • • 

&ln\ 

0 

1 

f>23 

’ • b 2 n 

'0 

0 

0 • 

■ ]/ 


in which the elements below the main diagonal are all zero. When the 
system (22-4) is written out in full, it has the appearance of Eqs. (4-2) in 
Chap. 9, whose solutions, as shown in Sec. 4, Chap. 9, can be obtained 
quite readily. 1 

Among other methods for solving the system (22-2) is the method of 
orthogonalization, the essence of which is as follows. Let us seek a matrix 
C such that the product 

CA - Z> 


1 This is the so-called Gauss reduction method discussed in Chap. 9. 


(22-5) 



SEC, 22] LINEAR VECTOR SPACES AND MATRICES 351 

is an orthogonal matrix. Since D is required to be orthogonal, it follows 
from (19-9) that 

DD r - D'D » /, (22-6) 

where I)' is the transpose of D. 

On multiplying (22-2) on the left by A'C'C, we get 

A'C'CAx = A'C'Cb, (22-7) 

and since 

A'C' - (CAY * D f 

by virtue of (17-11) and (22-5), we can write (22-7) as 

D'Dx « D'Cb. 

However, by (22-6) D'D = /, so that we finally have 

x - D'Cb. (22-8) 

Formula (22-8) gives the solution of the system (22-1) once a matrix C is 
determined. We do not present the classical procedure for constructing 
C (known as the Gram-Schmidt method) because of the rather special 
character of the problem. 




CHAPTER 5 


VECTOR FIELD THEORY 




Coordinates and Functions 

1. Curvilinear Coordinates 357 

2. Metric Coefficients 360 

3. Scalar and Vector Fields. Gradient 367 

4. Integration of Vector Functions 372 

5. Line Integrals Independent of the Path 378 

Transformation Theorems 

6. Simply Connected Regular Regions 382 

7. Divergence 384 

8. The Divergence Theorem 388 

9. Green’s Theorem. Line Integral in the Plane 391 

10. Curl of a Vector Field 396 

11. Stokes’s Theorem 400 

Illustrations and Applications 

12. Solenoidal and Irrotational Fields 402 

13. Gradient, Divergence, and Curl in Orthogonal Curvilinear 

Coordinates 405 

14. Conservative Force Fields 408 

15. Steady Flow of Fluids 411 

16. Equation of Heat Flow 414 

17. Equations of Hydrodynamics 416 


355 




This chapter is concerned with a study of scalar and vector functions 
defined in the familiar three-dimensional space. It includes a discussion 
of curvilinear coordinate systems and a derivation of several transforma- 
tion theorems involving line, surface, and volume integrals. These theo- 
rems, usually associated with the names of Gauss, Green, and Stokes, are 
indispensable in the study of mechanics of fluids, thermodynamics, and 
electrodynamics and in virtually every branch of mechanics of deformable 
media. 


COORDINATES AND FUNCTIONS 

1 . Curvilinear Coordinates. The chief advantage of formulating rela- 
tions among geometrical and physical quantities in the form of vector 
equations is that the relations so stated are valid in all coordinate systems. 
Only when one comes to consider a special problem involving numerical 
computations does it prove desirable to translate vector equations into 
the language of special coordinate systems that seem best adapted to the 
problem at hand. For example, in analyzing vibrations of clamped rec- 
tangular membranes, it is usually advantageous to express the displace- 
ment vector in cartesian coordinates. In the study of heat flow in a sphere, 
the geometry of the situation suggests the use of spherical coordinates, 
while problems concerned with the flow of currents in cylindrical con- 
ductors may indicate the use of cylindrical or bipolar coordinates. All 
these coordinate systems are but special cases of the general curvilinear 
coordinate system which we proceed to describe. 

Let us refer a given region R of space to a set of orthogonal cartesian 
axes y u y< h 2/3. We denote the coordinates of any point P in R by (2/1 ,2/2, 2/3) 
(Fig. 1) instead of the familiar labels (x,y,z). A set of functional relations 

*1 ** 

*2 « * 2 ( 2 / 1 , 2 / 2 , 2 / 3 ), 

*3 * * 3 (VU 2 / 2 , 2 / 3 ), 

357 


( 1 - 1 ) 



358 


VECTOR FIELD THEORY 


[chap. 5 

connecting tine variables y Xf y 2) 2/3 with three new variables x lt x 2) x 3 is 
said to represent a transformation of coordinates . We shall suppose that 
the functions Xi(yi f y 2 f yz) (i * 1 , 2 , 3 ) are single-valued and are continu- 
ously differentiable at all points of the region R and that Eqs. (1-1) can 
be solved for the 2/* to yield the inverse transformation 

Vi * Vi(*i,* 2 ,*s)f 

V 2 * 2 /a(*i,* 2 ,* 3 ), ( 1 - 2 ) 

VS - Vs(Xl>Z2j X 3)t 

in which the functions 2/,*(x 1 ,22**3) are single-valued and continuously 
differentiable with respect to the variables x t . The transformations ( 1 - 1 ) 



and ( 1 - 2 ) with these properties establish a one-to-one correspondence be- 
tween the triplets of values (2/1 ,2/2, 2/3) and (21,22**3)- We shall term the 
triplet of values (*i,* 2 ,2 3 ), corresponding to a given point P(t/i,2/2,2/3), 
the curvilinear coordinates of P, and shall say that Eqs. ( 1 - 1 ) define a 
curvilinear coordinate system xi,x 2 j x 3 . The reason for this terminology is 
the following: If we set in ( 1 - 1 ) x x = cy ( a constant), the equation 

*i(2/i, 2 / 2 , 2 / 3 ) * ci (1-3) 

represents a certain surface S x . Similarly, equations 

*2(2/1, 2/2, 2/3) * c 2 , ( 1 - 4 ) 

and *3(2/1, 2/2, 2/3) * c 3 , ( 1 - 5 ) 

represent surfaces S 2 and S3- These surfaces, shown in Fig. 2, intersect 
at the point P whose cartesian coordinates (2/1 ,3/2, 2/3) can be obtained by 
solving Eqs. ( 1 - 3 J to ( 1 - 5 ) for the y im 



COORDINATES AND FUNCTIONS 


359 


SEC. 1] 

The surfaces Si are called coordinate surfaces, and their intersections 
pair by pair are coordinate lines x lf x 2 , x 3 . Thus, the x x coordinate line 
is the line of intersection of the surfaces x 2 = c 2 and z 3 » c 3 . Along this 
line the only variable that changes is x lf since x 2 = c 2 and x 3 = c 3 along 
the line x x . Similarly, along the x 2 coordinate line the only variable that 
changes is x 2f while along the x 3 line the only variable that changes is x 3 . 

A very special case of the set of Eqs. (1-1) is 

*i * Vu 

*2 = 2/2, (1-6) 

£3 = 2/3- 

If we set x t *= Ci ( i = 1, 2, 3) in 
(1-6), we get three planes yi = c t 
perpendicular to the y coordinate 
axes. These planes intersect at the 
point (ci ,c 2 ,c 3 ) . The coordinate sur- 
faces in this case are planes, and 
their intersections pair by pair are 
straight lines parallel to the coordi- 
nate axes. 

As a more interesting example 
consider a transformation 

yi « r cos 0, 

2/2 = r sin 0, (1-7) 

2/3 = 2, 

which is of the form (1-2) if we set x x - r, x 2 = 0, x 3 = z. The inverse 
of (1-7) is 

r = 4- ^2/f + yl, 

Vo 

6 = tan""* 1 — * (1-8) 

2/i 

2 - 2/3, 

and it is single-valued if we take 0 < 0 < 2ir and r > 0. The surface 
r « ci is a circular cylinder y\ + y\ = c\ whose axis coincides with the 
y z axis (Fig. 3). The surface 0 = c 2 is the plane y 2 = (tan c 2 )y x containing 
the 2/3 axis, while the surface z « c 3 is the plane 2/3 = c 3 perpendicular 
to the 2/3 axis. The r, 0, and z coordinate lines are shown in Fig. 3, and 
we recognize that the curvilinear coordinate system r, 0, z is the familiar 
system of cylindrical coordinates. 




360 VECTOR FIELD THEORY 

As a final example, consider the transformation 
yi « p sin 6 cos 


with the inverse 


2/2 ** p sin $ sin <t>, 
2/3 - P cos 0, 


[chap. 5 


(1-9) 



0 = lan 


<t> ~ tan" 


+ j/3, 


( 1 - 10 ) 


2/2 

2/1 


which is single-valued if we suppose 
that p>0, O<0<7r, Q<<£< 2jr. 

The transformation defines a 
spherical system of coordinates. 
The coordinate surfaces p = const, 
6 - const, and <t> = const are, re- 
spectively, spheres, cones, and 
planes, shown in Fig. 4. The co- 
ordinate lines are the meridians, the lines of parallels, and the radial lines. 


PROBLEMS 

1- Discuss the curvilinear coordinates determined by 

V\ * 3*1 -f X2 -f *3, 

V2 « Xi - *2 + * 3 , 

2/8 * 2jj + X2 - i*3. 

2, Show by geometry that the coordinate lines in cylindrical and spherical coordinate 
systems intersect at right angles. 

2. Metric Coefficients. In this section we introduce an abridged no- 
tation which will enable us to write many formulas compactly and without 
loss of clarity. Thus, wc shall write the set of three equations of trans- 
formation (1-1) in the form 

x * * x t(y\>V 2 , 2 / 3 ), i = 1, 2, 3, (2-1) 

aAd their inverse (1-2) as 


Vi * V%(*i&&). 


(2-2) 


COORDINATES AND FUNCTIONS 


361 


SBC, 2] 

Throughout this section we shall suppose that the Latin indices i } j, k 
have the range of values 1, 2, 3. 

If P(yuy 2 ,ya) is any point referred to a set of cartesian axes y (Fig, 5), 
its position vector r can be written in the form 

X — i 1 y l + i 2 y 2 + hyz, (2-3) 



where the ij, i 2 , i 3 are the unit base vectors, which in Chap. 4 we denoted 
by i, j, k. 

The square of the element of arc ds along some curve C has the form 
(ds) 2 = (</?/] ) 2 + ( dy 2 ) 2 + (# 3 ) 2 , (2-4) 

and since 

dr = ii dyi + i 2 dy 2 + 13 dy 3) (2-5) 

we can write (2-4) as a scalar product 

3 

(ds) 2 = £ dy % dy x = dr-dr. (2-6) 

t ==1 

If we replace the y x in (2-3) by their values in terms of the xs with the 
aid of (2-2), r becomes a function of the variables x % and we can write 


dr 


dr 

dxi + 

dxi 

v dT * 

2- 

dx. 


dr 

dx 2 


dx 2 + 


dr 

-—dx 3 

dx 3 


dt(xi,x 2 ,xa) 


(2-7) 


Now, the symbol 



VECTOR FIELD THEORY 


362 


[chap. 5 


denotes the derivative of r with respect to a particular variable Xi (i » 
1, 2, 3) when the remaining variables are held fast. Thus, if we fix the 
variables x 2 and x 3 by setting x 2 = c 2 and x 3 » c 3 , r becomes a function 
of X\ alone, and hence the terminus of r is constrained to move along the 
Xi coordinate line in the x coordinate system determined by Eqs. (2-1)* 
Consequently, the vector 

dr Ar 

— = lim — 
dxi a*, o AX! 


is tangent to the coordinate line x x . Similarly, we conclude that the 
vectors dr/dx 2 and dr/dx 3 are tangent to the x 2 and x 3 coordinate lines, 
respectively (Fig. 5). If we denote these vectors by a„ so that 


we can write (2-7) as 


dr 



3 

dr = 23 a * dx,* 

*~i 


(2-8) 

(2-9) 


and hence Eq. (2-6) assumes the form 


(ds ) 2 - 


( 2 - 10 ) 


On expanding the scalar product in (2-10), we see that formula (2-10) 
can be written as 

3 3 

(ds) 2 = 23 23 • SLj dxi dxj 

*~1 ;-l 

and hence, with g x j defined by 


we can write it as 


35 Qn, 

3 3 


(ds) 2 « 23 13 Qij dx % dxy. 
i-i y-i 

In expanded form this reads 

(ds) 2 » 0ii (dxi) 2 + 012 dxi dx 2 + 0 i3 dx i dx 3 
+ 021 dx 2 dxi + 022(dx 2 ) 2 + 023 dx 2 dx 3 
+ 031 dx 3 dxi + 032 dx 3 dx 2 + gw(dx 3 ) 2 . 


( 2 - 11 ) 

(2-12) 


(2-13) 


Since a, • a ; = a ; *a l , we see from the definition (2-11) that g t j = Thus 
the quadratic differential form (2-13) is symmetric. 

For reasons which will appear presently, the coefficients g X j in this 
quadratic form are called metric coefficients. We shall see that they can 



SEC. 2] COORDINATES AND FUNCTIONS 363 

be computed directly from Eqs. (2-2) without first calculating the vectors 
a,-. 

The vectors a t , which were found to be tangent to the coordinate lines 
Xi at a given point P, are called base 
vectors in the curvilinear coordinate 
system x. Any vector A with the 
origin at P can be resolved into com- 
ponents A i, A 2 , A 3 along the direc- 
tions of the vectors fti, a 2 , a 3 (Fig. 

6). Thus, the base vectors a, play 
the same role in the system x as the 
base vectors i x , I2, is do in the car- 
tesian system y. It should be noted, 
however, that while the magnitudes 
and directions of cartesian base vec- 
tors are fixed, the vectors a„ in gener- 
al, vary from point to point in space. 

From the definition (2-11) we see 
on setting i = j » 1 that the length __ 

of &i is | ai| « V^n- Similarly, |a 2 | = V022 and |a 3 | * V033. These 
vectors are orthogonal if, and only if, 

012 * 021 = *a 2 = 0, 

03i = 0i3 ~ ai*a3 ■* 0, 

023 ~ 032 ~ a 2 *a 3 = 0 . 

A curvilinear coordinate system for which these relations hold is called 
orthogonal , and we note that in an orthogonal system the quadratic form 
(2-13) has the structure 

(ds ) 2 * gnidxi ) 2 + 0 22 (dx 2 ) 2 + 033(d^) 2 . (2-14) 

To get at the meaning of the coefficients g n , 022, and g 33 , we note that 
when an element of arc ds is directed along the x x coordinate line, dx 2 » 
dx 3 = 0, since along the x x line x 2 and x 3 do not vary. Thus, (2-14) gives 
in this case 

(d«i) 2 « 0n(dx x ) 2 , 

so that ds x « V^gndx x . (2-15) 

Thus, the length of the arc element ds x along thejr x coordinate line is 
obtained by multiplying the differential of x x by *\/0ii- Similarly we find 
that the differentials of arc ds, along the x 2 and z 3 coordinate lines are 

ds 2 m dx 2 , ds 3 ■* dx a . (2-16) 




364 


VECTOR FIELD THEORY 


{CHAP, 6 

Since the ds,- and the dxi are real, we conclude that gu >0 f 022 > 0, 
033 > 0. In orthogonal cartesian coordinates (ds) 2 is given by the formula 
(2-4), and hence in such a system gu *= 022 = 033 353 h 
An element of volume dr in general curvilinear coordinates is defined 
as the volume of the parallelepiped 

dr = |ax*a 2 x a 3 1 dx x dx 2 dx s (2-17) 

constructed on the base vectors a,-. If the system is orthogonal, (2-17) 
reduces to 

dr = VV a6 r 23S33 dr; dx 2 dx 3 , (2-18) 

as is immediately obvious from (2-15) and (2-10). 

When a curvilinear coordinate system x is determined by equations of 
the form (2-1), we can write the inverse transformation (2-2) as 


Vk = yk{x l} x 2l xz) (2-19) 

and deduce the metric coefficients g l} as follows: On differentiating Eqs. 
(2-19) with respect to x % we get 


But in cartesian coordinates 


dy k = X) “ dz x . 

j -1 OXi 


ds 2 = X dy k dyk, 

k^i 


and the substitution from (2-20) in this formula yields 1 

3 r ^ 3 


d* 2 - z fr— d*] 

£i L,-l "l toy ’ J 

I «=: 1 1 '£ = 1 dXj y 

On comparing (2-21) with (2-12), we see that 

^dy k dy k . , , „ „ 

— M = 1,2,3. 

*-l t)X,- toy 


( 2 - 20 ) 


( 2 - 21 ) 


( 2 - 22 ) 


This is the desired formula for the calculation of metric coefficients. 

To illustrate the use of (2-22) consider a coordinate system defined by 
Eqs. (1-7), which we write in the form 

1 Note that the summation index can be changed at will so that 



SEC, 2 1 


COOHMNATES AND FUNCTIONS 


365 


Vl m x x cos x 2 , 
y 2 * x x sin x 2 ) 

2/a 3=1 *3, 

to agree with the notation used in this section. From (2-22) we have 

-©+©+©■ 

~ cos 2 x 2 + sin 2 x 2 + 0 = 1 , 


Qn 


*-©■+©+(£ )’ 

= xf sm 2 x 2 + *1 cos x| + 0 = x\, 


933 — 


© ,+ ©’ + ©' 

\dXs/ \dxs/ \dx 3 / 


0 + 0+1 


1, 


- f ^2 dj /2 ^ d //3 
d:rj dx 2 dxi dx 2 dxi dx 2 

- cos x 2 ( —Xx sin x 2 ) + sin x 2 (x { cos x 2 ) + 0 » 0. 

We find in tiie same way that g 2 :\ = (j\n = 0. Hence the system under 
consideration is orthogonal. The expression for ds 2 is 

3 3 

* 2 = S n g„ dx, dx, 

.= 1 ;~1 

= (dxi) 2 + x 2 i(dx 2 ) 2 + (fbr 3 ) 2 , 

which is a familiar formula for the square of the arc element in cylindrical 
coordinates if we recall that x x = r, x 2 = 0, :r 3 — 2 . Since this system is 
orthogonal, the element of volume is given by (2-18), which in our case 

dr = rdrdSdz. 

Example ; Obtain expressions for the elements of arc and volume in the coordinate 
system x defined by 

yx » x\ + X2 + a*, 

j/2 » xi - 12 - £3, (2-23) 

2/3 ** 2#1 +2*2 - X8, 

and discuss the system. 

On making use of formula (2-22) we find as in the preceding illustration that 
011 ** fi , 022 3 , 088 ** 3 , 012 *• 2 , 028 ** l f 018 “* -“ 2 . 



366 

Hence 


VECTOR FIELD THEORY 


[CHAP. 5 


ds* m 6(dxj )* 4 4 dx\ dx% — 4 dx\ dx$ 4 2 dxi dz% 4 3 (dxa)* 4* 3 (dz^) 2 . 

The system is dearly not orthogonal, and to compute dr we shall make use of formula 
(2-17). Now 

r «* IiFi 4 ky* 4 lays 

** ii(xi 4* x* 4- x$) 4- 4(xi — x* — £ 3 ) 4* 4 ( 2 xi 4* x* — x«) 
and hence the base vectors a« — dr/dxi are 

*1 “ U 4- h 4- 2ia, 

a 2 « ij — i 2 4 - i 3 , 


Thus, 

dt 


«s « ii — is — 4. 


|Ai*A 2 x As \ dx\ dX2 d>Xz 


1 

1 

1 


1 

-1 

-1 


2 

1 

-1 


dz\ dx% dx% 


4 dx 1 dx 2 dz$. 


On solving (2-23) for the x, we get 

xi « Hyi 4- Mv 2 , 

x 2 - -J4yi - Mv* 4 Hi/s, 

x 3 * %yi 4 J^j /2 ~ V 2 V a- 

The coordinate surfaces x» » c* are planes, and the coordinate lines Xi are therefore 
straight lines. 

The system in this example is a special case of an affine coordinate system 
determined by the transformation 


Vi — cixiXi 4- a i2 x 2 4* i » 1, 2, 3, (2-24) 

in which the a t y are constants. Affine transformations (2-24) occur in the 
study of elastic deformations, in dynamics of rigid bodies, and in many 
other branches of mathematical physics. 


PROBLEMS 

1. Discuss in the manner of the preceding example a coordinate system x determined by 

1 1 

w " Z^ Xl " V5**- 



BEC. 3] COORDINATES AND FUNCTIONS 367 

2. Compute the metric coefficients appropriate to a spherical coordinate system defined 
by Eqs, (1-9), and thus show that 

(d*)* - (dp) 2 + p 2 (dS) 2 + p 2 sin 2 0 (d 4 >)* 

and dr ■» p 2 sin 6 dp d$ d<t>. 

3. If R — ix 4- jy 4* k* is the position vector of a moving point P(z,y,z) in cartesian 
coordinates, show that the unit base vectors e r , e$, e* in cylindrical coordinates (r,$ t z) 
[see (1-7)1 are 

e r «■ i cos $ 4* j sin 0, e# — i sin 0 4- j cos e, * k. 

Show that R *• re r 4* ze gt compute dR/dt and d*Rfdt 2 , and thus show that the velocity 
v and the acceleration a of the point P are 



4. If R «• ir 4- jy 4* kz is the position vector of P(x t y,z ) in cartesian coordinates, 
show that the unit base vectors e p , e*, e* in spherical coordinates defined by Eqs. (1-9) are 

e p « i sin 0 cos <f> 4- j sin 0 sin <t> 4- k cos 0, 

©$ i cos 0 cos <f> 4* j cos 6 sin <jb — k sin 0, 

e* — i sin <*> 4- j cos <t>. 

6. If the position vector R of a moving point P in spherical coordinates is written as 
R ® pe p , where e p is the unit vector in the direction of the increasing coordinate p, use 
the results of Prob. 4 to show that 

dR dp d$ d<t> 

3. Scalar and Vector Fields. Gradient If in some region of space a 
scalar u(P) is defined at every point, we say that u(P) is a scalar point 
function. An example of such a function is the temperature at any point 
in a solid. A function v(P) defining a vector at every point P of the 
given region is a vector point function. An example of vector point function 
is the velocity at any point P of a fluid. The regions of definition of 
scalar and vector point functions are sometimes called fields, and one thus 
speaks of scalar and vector fields. Unless otherwise noted, we shall assume 
that w(P) and v(P) are single-valued functions. 

To facilitate calculations involving scalar and vector point functions, it 
is often convenient to refer the region of their definition to a special co- 
ordinate system x. If this is done, the coordinates of P can be denoted by 
{x\,x 2l x z ) and u(P) and v(P) can be denoted by u(x u t 2 ,xz) and v{x u x 2 ,x z ), 
respectively. As explained in the preceding section, v(x h x 2t x z ) can then 
be represented in terms of its components v t (xi,x 2 , x z ) (i = 1, 2, 3) along 



VECTOR FIELD THEORY 


[CHAP. 5 

the appropriate base vectors at It should be noted, however, 

that the introduction of coordinate systems is a matter of convenience 
and that u(P ) and v(P) depend only on the choice of P in the field and 
not on any special reference frame selected to locate P. The fact that 
scalar and vector point functions are independent of coordinate systems 
is spoken of as invariance , and we shall see that it is possible to associate 
with u(P) and v(jP) certain new scalar and vector functions which have 
important invariant significance. 

We say that u(P) and v(P) are continuous at P if 

lim u(P') ** u(P) and lim v(P') « v(P) 

p’ -+ p p' p 

for every choice of P l in the neighborhood of P. Functions continuous 
at every point of the region are said to be continuous in the region. 




Fra. 7 Fig. 8 


Let u(P) be a continuous scalar function in the given region. Wo 

- > 

select a point 0 in this region for the origin of position vectors r ss OP 

— -> 

If P r is some point in the neighborhood of P, we denote 0P f by r' and write 

r-r + Ar. 

The difference quotient 


tt(P') - u(P ) 
|Ar| 


u{P') - u{P) 
As 


(3-1) 


where |Ar| * As f gives an approximate space rate of change of u(P), and 
w r e can study the limit of (3-1) as P f is made to approach P along the 
rectilinear path Ar. If this limit exists, we shall write 

u(P') - u(P) du 

lim = — * 

Am ~+ 0 AS d8 


(3-2) 


COORDINATES AND FUNCTIONS 


SEC. 3] 


369 


and call it the directional derivative of u(P) in the direction specified by Ar. 
A different choice of P f yields a different vector Ar and in general a different 
value for du/de at P. 

A set of points for which u(P) has a constant value c determines a sur- 
face S called a level surface; we assume that at each point of S there is a 
uniquely determined tangent plane. Let us consider a pair of such sur- 
faces S and S' determined by u » c and u = c + Ac, where Ac is a small 
change in c (Fig. 8). If P is a point on S and P* on S', the change A u ss 
u(P f ) — u(P) is Ac, and this is independent of the position of P* on S'. 
But the average space rate of change 


u(P') — u(P) A u 

j Ar j ” As 


(3-3) 


clearly depends on the magnitude of Ar, The limit of this ratio as Ar is 
made to approach zero by making Ac — » 0 is the directional derivative 

(3-2) in the fixed direction determined by Ar. The greatest space rate 

— ■> 

of change of u will occur when P' is taken on the normal PQ as An to the 
surface S (Fig. 8), since for this position of P' the denominator (Ar| in 
(3-3) is not greater than | An | . Indeed, 


An ~ Ar cos B, 


where B is the angle between the normal PQ to 5 and PP 
On taking account of (3-4), we conclude that 


(3-4) 


du 1 du du 

— — H5 — sec Q. 

dn cos B ds ds 


(3-5) 


The derivative du/dn in the directs Jhe normal to the level surface 
u = const is called the normal derivative oj u(P). 

If n is a unit vector at P, pointing in the direction for which Au > 0, 
we can construct a vector, called the gradient of u , namely, 

du 

grad u SB n — (3-6) 

dn 


This vector represents in both the direction and magnitude the greatest 
space rate of increase of w(P), provided, of course, that du/dn 0. The 
gradient vector (3-6) is clearly independent of the choice of coordinate 
systems and hence is an invariant. If we introduce the familiar cartesian 
coordinates xyz and denote u(P) by w(:c,y,z), then, as in Chap. 3, Sec. 8, 

du du dx du dy du dz 
ds dx ds dy ds dz ds 


(3*7) 



870 


VECTOR FIELD THEORY 


[CHAP. 5 


where dx/ds = cos (#,$), dy/ds * cos (y,s), dz/ds = cos (z,s) are the direc- 
tion cosines of the unit vector s in the direction of the arc element ds 
(Fig. 9). In this case the position vector r or P is 


and 


ix + )y + kz 
dr 
ds 


. dx dy dz 

i X + i T + k T' 

ds ds ds 


(3-8) 



of s coincides with that of the normal 
conclude that 


We see that (3-7) can be written as 
the scalar product of the vector 

du du du 


Vus i h j t-k — 

dx dy dz 

and the unit vector s in (3-8). 


(3-9) 

Thus, 


— - VW'S. (3-10) 

ds 

Inasmuch as the greatest value of 
du/ds is assumed when the direction 
n to the level surface u * const, we 


Vu = grad u, 


(3-11) 


for the right-hand member of (3-10) can be interpreted as the component 
of the vector Vu in the direction s and the maximum component du/ds 
is obtained when s is directed along Vu. 

It follows from (3-9) and (3-11) that a formula for calculating grad u in 
cartesian coordinates is 

du du du 

grad u = i hj b k — (3-12) 

dx dy dz 


On comparing (3-6) and (3-12), we see that 

du 

lgradu| * — 
an 


j/du\ 2 /du\ 2 ( du \ 2 

VU + U' + U 


Formula (3-9) suggests a definition of the differential vector operator 
V, called del or ndbla , 

d d 0 

V=i- -fj — + k— . (3-13) 

dx dy dz 


analogous to the scalar differential operator D introduced in Chap. 1, 
Sec. 23. The product of V and the scalar u(x,y,t) is interpreted to mean 



371 


SEC* 3] COORDINATES AND FUNCTIONS 

(3-9). The reader will show that 

V(u + v) ** Vu + Vv, 

V(uv) ® uVv + vVu , 

whenever u and v are scalar functions of ix^y^z). A formula for grad u 
in orthogonal curvilinear coordinates is deduced in Sec. 13. 

The directional derivative dv/d$ of a vector point function v(P) is 
defined by formula (3-2) in which u(P) is replaced by v(P). When v(P) 
is expressed in the form 

v ** iv x 4" K + kv z , (3-15) 

where 1, j, k are the base vectors in the system x,y,z , 


dv 

ds 



dv v 

+ i~ + k 
ds 


dv t 

ds 


(3-16) 


We have already employed a similar formula in Chap. 4, Sec. 7, to cal- 
culate the derivatives of the position vector R = ix + jy + kz with respect 
to the time parameter t. 


Example 1. Find the directional derivative of u — xyz* at (1,0,3) in the direction of 
the vector i — j 4- k. Compute the greatest rate of change of u and the direction of 
the maximum rate of increase of u. 

On substituting u m xyz 2 in (3-9), we find that the gradient u is given by 


At (1,0,3) 


Vu «* iyz 2 ~f jrz 2 -f k2xyz. 
Vu « iO + j9 + k0 - 9j. 


Thus, the greatest rate of change | Vu | *9, and the direction of the maximum rate of 
change is along the y axis. Since the unit vector s in the direction of the vector i — j -b k 
18 


6 


1 

\/5 


0 - J + k). 


we find on using (3-10) that the desired directional derivative is 

du 1 9 

— - Vu-s - 9j • - j + k) - - 

Example 2. Find the unit normals to the surface x 2 — y 2 -j- z 2 - 6 at (1,2,3). 

The surface in this example is a level surface for the function u ■* x 1 — y % -f s*. 
Since the gradient of u is normal to the level surface u ** const, we have by (341) 

grad u «* Vu «« i2x — \2y + k2z. 


which at (1,2,3) has the value 


Vu ~ i2 - j4 + kfi. 



VECTOR FIELD THEORY 


872 


[chap, S 


But this vector Is directed along the unit normal n to u « x 2 — y % + ** *■ 6 in the direc- 
tion of increasing u. Hence 


Vu 

n mt — 

I VI* | 


1 

V5$ 


(12 ~ J4 + 2*6). 


The direction of the other unit normal vector is opposite to this. 


PROBLEMS 

1. Compute the directional derivative of u * x 2 4* y 2 4* * a at (1,2,3) in the direction 
of the line 

x y t 
3 " 4 ~ 5* 

Find the maximum rate of increase of u at (1,2,3); at (0,1,2). 

2. Find grad u if (a) u » (x 2 4* V 2 + z 2 )~ H t ( b ) u * log (x 2 -f y 2 + * 2 ). 

8. Find the directional derivative of u * x 2 y — y 2 z — xyz at (1,— 1,0) in the direc- 
tion of the vector i — j -f 2k. 

4. Find the directional derivative of u » xyz at (1,2,3) in the direction from (1,2,3) 
to (1,-1, -3). 

5. Find the unit normal vector in the direction of the exterior normal to the surface 
x 2 +2y 2 4-z 2 - 7 at (1, — 1,2). 

6. Find the unit vectors normal to xyz » 2 at (1, — 1, —2). 

7. Show that Vr n *» rar n " 2 r, where r « ix 4* jy + kx and r ** |r|. 

8. Use the result of Prob. 7 to compute the directional derivative of u = (x 2 4- y 2 + z 2 ) s 
at ( — 1,1,2) in the direction of the vector i — 2j k. 

9. Compute the directional derivative of 

v « i(x 2 - y 2 ) + j (xyz - 1) + k* 
at (1*2,0) in the direction from (1,2,0) to (0,0,0). 

4. Integration of Vector Functions. Integrals of vector functions with 
the integrands consisting of scalar products of vectors are defined in the 
usual manner. Thus if v(P) is a continuous vector point function specified 
along a curve C joining a pair of points P 0 , P' } and if t{P) is the position 
vector of P on C, then the integral 



is defined as the limit of a sum constructed as follows. Let (7, which we 
suppose to be sectionally smooth, 1 be divided into n arc elements As,- 
by inserting the points Pi (Fig. 10). We form the sum 

2>(P,)-Ar ( , (4-2) 

1 

1 This means that C consists of a finite number of segments with continuously chang- 

ing tangents. The toxm piecewise smooth is also used. 



COORDINATES AND FUNCTIONS 


373 


SRC. 4] 


where Ar* = r* + i — r t *, and compute the limit of this sum as n — » oo 
and every | Ar*| — > 0. The continuity of v(P) and the smoothness of C 
suffice to show that the limit of (4-2) 
exists, and we define the line integral 
(4-1) to be this limit. 

If v(P) is defined in some region 
containing several paths joining Po 
and P, then the integral (4-1) will 
ordinarily have different values when 
computed along different paths. In 
exceptional circumstances, discuased 
in the following sections, these values 
may turn out to be equal. 

If we introduce the xyz coordinate 
system and write 

v(P) = v{x,y,z) S w x (x,y,z) -f j v v (x,y,z) + kv z (x,y,z), 
dr see i dx + j dy + k dz, 
the integral (4-1) becomes 

p f 

, Mx,y,z) dx + v v (x,y,z) dy + v *(x,y,z) dz]. (4-3) 

0 

When the equations of C are given in parametric form 


P'mP, 



X — x(t), 

y = y(t), 

z = z(l) K 


to <t <t‘ 


(4-4) 


where the values to, t' of the parameter t correspond to the end points P 0, P r 

rt ' 

of C, the integral (4-3) can be expressed as a definite integral 1 / F(i) dt and 

JtQ 

evaluated by the usual means. 

Similarly, the surface integral 

f v*n da (4-5) 


where n is a unit normal specified at all points of a sufficiently smooth 
surface 2 2, can be defined as the limit of the sum 

1 See the examples at the end of this section. The equivalence of the integral (4-3) 
and the ordinary Riemann integral, when (4-4) holds, is easily seen by comparing the 
sums of which these integrals are the respective limits. 

* W© assume that the surface 2 is two-sided and that n is directed toward one side. 
This normal we elect to call positive. If the surface is closed, it is customary to regard 
the exterior normal as positive. 



374 


VECTOR FIELD THEORY 


[CHAP. S 


k 

lim £ v(P,)-n(P,) (4-6) 

* * i-1 

In constructing this sum it is supposed that the surface 2 is divided into 
k elements of areas Acr* and P, is chosen somewhere in the element A<r x . 
The limit is then computed by increasing the number k of elements in 
such a way that the maximum diameter of every Acr,- approaches zero. 


Formally one is tempted to extend these “limits of the sum" definitions to such 
symbols as 

*(/>)*, f z <P)<b, l HP) dr, (4-7) 


in which v(P) is a vector function and ds, da, and dr, respectively, are the elements of 
arc length, surface, and volume. Thus, there is a temptation to define the volume 

integral jfv(P) dr by the formula 

r * 

/ v(P) dr « lim £ v(P.) Ari, (4-8) 

Jr k~* « t_l 

in which it is imagined that the volume r is divided into elements of volume At*. A 

k 

definition such as (4-8) requires forming sums X) v(P*) At* of the bound vectors v(F\) 

t-i 

which are determined at different points of the body. There is a question if the rules 
for addition of free vectors given in Chap. 4, Sec. 2, can be used to provide a sensible 
definition of (4-8). Without going into details we state it as a fact that the definition 
(4-8) makes sense in those geometries where the distance between a pair of points is 
given by the Pythagorean formula. 1 

If v(P) is expressed in terms of its cartesian components as v » iv z (x,y,z) -f }v v (x t y,z) *f 
k v»(x,y,z), the integrals in (4-7) can be reduced to the evaluation of three ordinary inte- 
grals by writing, for example, 

j V(P) dr » ij V z dr *f )J Vydr + kj v,dr. 

No such simple means of evaluating integrals of the type (4-7) are available in curvilinear 
coordinates because the base vectors in curvilinear coordinate systems vary from point 
to point in space. This remark may serve to explain why cartesian coordinates are so 
prominent in calculations involving vectors, 
line integrals of the form 

f [P(x,y,z) dx + Q(x,y,z) dy 4* R(x,y t z) dz] t (4-9) 

Jc 


which is identical with (4-3), are frequently defined without reference to vectors, but as 
we shall see, the definition adopted here has many interesting and immediate physical 
interpretations. 

1 Spaces so metrised are called Euclidean, and it is only with such that we are con- 
cerned in this book. 



375 


BBC. 4 } COORDINATES AND FUNCTIONS 

Example 1 . Evaluate the integral / t*dx when C is the helical path 

Jc 


x •» cost, 

V ~ aia (4-10) 


* -*» 

joining the points determined by t « 0 and f « r/2 and also when C is the straight line 
joining these points. 

Since r ** ix -f ft/ -f* k*, we get, on using (4kl0), 

r « i cos t -f j sin t -f k*, 
dr «* ( — i sin t -f j cos t -f~ k) dt. 
r r* 12 w* 

Hence / r*dr « / fdf «» — • (4-13) 

Jc Jo 8 

If the path C is a straight line joining the same points (1,0,0) and (0,l,ir/2), we can 
write its equation in vector form as 

r * n *f (r* - riX (4-12) 

where ri and ra are the position vectors of (1,0,0) and (0,1 , t/ 2), respectively (Fig. 11). 



The parameter t clearly varies between 0 and 1, since for t — 0, (4-12) yields r ■■ n 
and for t «« 1, r * r 2 . But n ■■ i, r* «- j + (x/2)k, so that (4-12) reduces to 

r “ i + ( j+ i k " 1 ) < - 

Hence / c r-dr “ £ [ + ( 2 + j) *] * “ J* 

This is the same value as we got for the helical path. In the following section we shall 
•ee why this particular integral is independent of the path. 



VECTOR HELD THEORY [CHAP. 5 

Example 2. Compute the value of / v»dr, where v *■ iy -f j2ar and C is the straight 

Jc 

line joining (0,1) and (1,0). Discuss also when C is the arc of a circle centered at the 
origin. 

* ( JSkice r ** ix -f jy, we have dr « i dx + j dy and therefore 

f v 'dr * f (y dx -f 2x dy). (4-13) 

* rJ ,fc > Jc Jc 

To evaluate this integral along the rectilinear path in Fig. 12, we write the equation of 
the path in the form 

y «* _J_ 1 (4-14) 

and insert (4-14) in (4-13). Since dy » -dx, we get 

f vdr * f [(— x 4* 1) dx — 2x dx] * f (1 — 3x) dx — — H. 

1 ' t Jc Jo Jo 


, yhe 1 iptegr%tioa here is performed so that the path C is traced from the point (0,1) to 



To compute the value of the integr al (4-1 3) over the circular path C f joining the 
same two points, we note that y ** y/\ — x % along C", dy * — x dx/y / 1 — x a , so that 



■L 


1 — x 2 dx 


2x 2 


1 1 -3x* 
y/l z 2 


dx 


VT 

w 

~ 4* 




Again, the path C" is traced out from (0,1) to (1,0). If the direction of description of 
C’ is reversed, so that the circle is traced out from (1,0) to (0,1), the limits in the inte- 
gral must be interchanged and we get -f t/ 4 for the value of the integral, 

n Example 3. Evaluate / v*dr, where v « (iy — jx)/(x 2 -f y 2 ) and C is the circular 

P i it ft v , Jc 

path** + p* •* 1 described counterclockwise. 



COORDINATES AND JUNCTIONS 


SEC. 4] 


377 


This integral c an be evaluated as in the preceding example by substituting in the 
integrand y ** s/\ — x % for points on the upper half of the circle O' and 
on the lower half. It is simpler, however, to write the equations of the path in para- 
metric form , rf iu 


We thus get 


X ** cos 0, ) 

} 0 < e < 2t. 

y «* sin 0 ,J 


t*,. 


(4-15) 

V*'Oi i*U) k ’i V * r L 


t •* ix •+■ jy ■* i cos $ 4* j sin & f 


dr « ( — i sin 0 + j cos 8) dd$ 


v 


i sin 0 — j cos $ 
sin 2 0 + cos 2 0 


i sin 8 — j cos 0. 


h:') k/'J 
2 , ' * ti 


Hence 




sin 2 0 — cos 2 6) de 


~2 v. 


If the path is traced in the clockwise direction, we get +2*-. 

It may prove instructive to evaluate this integral over the square C' formed by the 
lines x « ±1, y * rhl (Fig. 13). t 

The integral over C" is equal to the sum of four integrals evaluated otfCr the paths 
PQ, QR , RS, SP. 

Now along PQ, y * —1, dy «* 0, r ** Lr — j, dr * i dx, and v «« ( — i — ji)/(^ # + 1)’. 


Hence 



tan“ 


l 

M 

-1 


X 

2 * 


Along the path QR, x « 1, r « i + jy, dr — j dy, v « (iy — j)/(l -f y*), so that 


J v-dr « ( 
qr J - 1 


l 4- y 2 


In a similar way we find that 


"pul Sum 


% j • 


/ v*dr *» / v*dr — — 

•Jfls -w 2 

so that the integral 

2- ■ .! 

This time we obtained the same result as we did for the circular path:. 'In? Seel 15 we shaft 
see that this is not an accident and that the value of this integral for enfeqr 
enclosing the origin is — 2*\ 


PROBLEMS 


1, Evaluate the integral in Example 2 over the path C consisting o( straight-line 

segments joining the points (0,1), (0,0), (1,0) in that order. ' *■' ' ^ " 

2. Evaluate the integral in Example 1 over the polygonal path joining the points 
(1,0,0), (1,1,0), (1,1, x/2) in that order. 



378 


VECTOB HELD THBOBY 


[CHAP. 5 

8. Compute the value of the integral Jj(xy dx — ydy + dz) over the following paths: 

(a) Straight line joining (0,0,0) (1,1,1), 

(b) Straight line joining (0,0,1) and (0,1,1), 

(c) Straight line joining (0,0,0) and (1,2,3). 

Note that this integral has the form v-dr. 

4u Compute the integral ^ v-dr where v «* lx — jy 4- kz over the helical path in 

Example 1. Also evaluate it over the rectilinear path. 

5. Compute the work W done in displacing a particle of unit mass in a constant 
gravitational field F « -kg' along the following paths: 

(a) Straight line joining (0,0,0) and (1,1,1), 

(b) A polygonal path joining(0, 0,0), (1,1,0), (1,1,1) in thatorder. Hint:W » /V*dr. 

JQ 

8. line Integrals Independent of the Path. A special case of line integral 

f c vdt ~f c [v x (x,y,z) dx + V v (x,y,z) dy + v,(x,y,z) dz], (5-1 ) 

in which v(x,y f z) is known to be the gradient of some single-valued scalar 
specified in the region R containing C, frequently appears in ap- 
plications. Now, if v « Vu, then 

du du du 

v*dr — Vu«dr = — dx -| dy H dz 

dx dy dz 

m du } (5-2) 

and thus the integrand in (5-1) is an exact differential. We can, therefore, 
write 

f c v ' dT = fp 0 du=s ~ u ( p o), (5-3) 

where Po and P are the end points of the path C. 

This result is unique since u f by hypothesis, is single-valued. More- 
over, since it depends only on the end points P 0 and P, we see that the 
value of the integral in (5-3) is independent of the path joining these points. 
If Ci and C 2 are two different paths shown in Fig. 14, then 

/ Vu*dt ** f Vu-dr. ( 5 - 4 ) 

Jp o JPq 

Ci c t 


But along C 3> 



SBC. 51 COORDINATES AND FUNCTIONS 379 

and we can therefore write (5-4) as 

rP rP« 

/ Vu-dx + / Vu dt ** 0 

JPt Jp 

Ci c, 

or fvudx = 0, (5-5) 

where C is the closed path formed by Ci and <? 2 . 



The results embodied in (5-3) and (5-5) can be stated as a theorem. 

Theorem I. The hue integral \c Vu-dx, in which u is a single-valued 

continuously differentiable function in a given region R, is independent of the 
path , and hence it vanishes for every closed path drawn in R. 

At first glance Theorem I appears to contradict the result in Example 3 

of Sec. 4, where the integral J c v-dr with v = (iy — jx)/(x 2 + y 2 ) was 

considered. It is easy to check that v = — V tan”” 1 ( y/x), so that in this 
case u = — tan -1 (y/x). This integral does not vanish when evaluated 
over any closed path including the origin because the function tan~ l (y/x) 
is multiple- valued. Also, the continuity requirement of the theorem is 
not fulfilled by v » Vu at (0,0). 

We can also establish another important theorem which is a converse 
of Theorem L 

Theorem II. If a vector point function v is continuous in a given region R } 

and if the integral j c v*dr is independent of the path, then a single-valued 

scalar u exists such that v = Vu in R, 

We shall prove this theorem by actual construction of the function 
u(x,y,z) fulfilling the conditions of this theorem. 



VECTOR FIELD THEORY 


380 VECTOR FIELD THEORY [CHAP. 5 

By hypothesis, the integral J c v-dr when evaluated over any curve 

C joining Po(xo,yo,Zo) with P(x,y,z ) is independent of the path and thus 
defines a single-valued function 

u(x,y,z) = / (v x dx 4 - v v dy + v, dz). (6-6) 

^(aco,j/o.*o) 

We shall show that this function is, indeed, such that v = Vu. 

On replacing x by x + Ax in (5-6), we get 

r(x+Ax, y % z) 

u(x 4- Ax, y, z) » / ( v x dx + v v dy + v 9 dz) (5-7) 

J(XQ,VQ,Z0) 

and on subtracting (5-6) from (5-7), we obtain 

r(x+Ax, y, z) 

u(x + Ax, y f z) — u(x,y,z) = / (t>* dx + v u dy + v 8 dz). (5-8) 

J(x,y.z) 

The integral in (5-8) is independent of the path joining (x,y,z) with 
(x + Ax, y, z ), and it suits our purposes to evaluate it over the rectilinear 
path y = const, z « const. Over such a path dy « dz « 0, and hence 
(5-8) yields 

rx-t-Ax 

u(x + Ax, y, z) - u(x,y,z) = / v x (x,y,z) dx. (5-9) 

But by the mean-value theorem for integrals 

J «x(x,y,z) dx = v x (£,y,z) Ax (5-10) 


where a; < £ < a: 4- Ax. The substitution from (5-10) in (5-9), on dividing 
by Ax, gives 

u(x + Ax, y, z) - u(x,y,z) 

' = v x ((,y,z). 

Ax 


Now, if we let Ax — * 0, we get 


du 

— “ v x{x,y,z) 

dx 

(5-11) 

by recalling the definition of partial derivative and by the fact that v x 
is continuous. In a similar way we prove that 

du 

— = v v {x,y,z) 
dy 

(5-12) 

du 

and — = v t {x,y,z). 

dz 

(5-13) 



SUC. 5 ] COORDINATES AND FUNCTIONS 381 

But the statements (5-11) to (5-13) are equivalent to the vector equation 
Vu « v, and the theorem is thus proved. 

It should be carefully noted that the key hypothesis which ensured the 
existence of a single-valued function u such that v « Vu is that the integral 

j c v*dr is independent of the path. The integrand v*dr « v x dx + v v dy 
+ v t dz may be an exact differential of a multivalued function u f in which 
case the integral f c V'dx may depend on the path. 

A differential form 


t; x (x,y,z) dx + v v (x,y,z) dy + t>*(x,y,z) dz, (5-14) 

in which v X) v y , v t are continuously differentiable single-valued functions, 
is said to be exact if 

du du du 

v x dx + v v dy + v t dz » — dx -j dy H dz, (5-15) 

dx dy dz 

where u is not necessarily single-valued. We can deduce a set of necessary 
conditions for (5-14) to be an exact differential as follows: If there exists a 
function u{x,y,z) such that (5-15) is true, then on setting x ~ const, y = 
const, z = const, in turn, we get 


( 516 } 


du du du 

v x 3 as f Vy = » Vg = 

dx dy dz 

Differentiating the first of Eqs. (5-16) with respect to y and the second 
with respect to x, we get 

dv z d 2 u dv v d 2 u 
dy dx dy dx dy dx 

But the mixed partial derivatives in these expressions are equal, since 
dv x /dy and dv v /dx are continuous by hypothesis (see Sec. 2, Chap. 3). 
Thus 


dv x dv y 
dy dx 

In a similar way we obtain two more relations 
dv y dv z dv e dv x 

dz dy dx dz 


(5-17) 


(5-18) 


The relations (5-17) and (5-18) give a necessary condition to be satisfied 
by the functions v X) v Vf v t in (5-15) if that differential form is to be an exact 
differential of some function u(x,y, 2 ). We shall see in Sec. 12 that these 
conditions suffice to ensure the existence of a function u such that (5-14) 
is equal to du. However, the conditions (5-17) and (5-18) do not guarantee 



382 


VECTOR HELD THEORY 


[CHAP. 5 

that u is single-valued. The question naturally arises : What supplementary 
conditions must be adjoined to Eqs. (5-17) and (5-18) to ensure that u 
is single-valued? A complete answer to this question is complex because 
it depends not only on the differentiability properties of v X} v V} v t but also 
on the geometry of the region in which these functions are defined. If 
the region of definition of these functions is simply connected and suf- 
ficiently regular to permit the use of certain integral transformation theo- 
rems discussed in Secs. 9 to 11, then u(x,y,z) determined from the formula 

u(x,y } z) ® / (v x dx + v v dy + v z dz) (5-19) 

J(ZQ,VQ>tQ) 

is single-valued. We describe these restrictions on the character of the 
region in the following section. 


PROBLEMS 

1. Show that the integral r*dr is independent of the path and find its value when 

computed over the rectilinear path joining (0,0,0) and (1,1,1). Hint: r*dr dO^r 2 ). 

2. Show that ( y — x 2 ) dx -f- (x 4* y 2 ) dy is an exact differential du, and, find u(x,y). 

3. Show that the conditions (5-17) and (5-18) for an exact differential can be written 
in symmetric form as 

i j k 

A A A 

dx dy dz 
Vx Vy V x 

4 . (a) Is yzdx zx dy xy dz an exact differential du ? If so, find u(x,y t z). (b) 

Evaluate the integral (yzdx -f zx dy -f- xy dz) over the rectilinear path joining 

(0,0,0) to a fixed point ( x t y,z ). 

x y f 

5. If v «* i — -f j -r , show that / v*dr * 0 for every closed path that 

x 2 + y 2 x 2 -f* y 2 Jc 

does not include the origin. What is the value of this integral over the circular path 


V x v 


y 2 =* 1? Find u such that du Vu*dr. 

6. 

7 . Find a function u such that 
x 


y — — JL i JL 1UU M/ OUVU UMr V U » 

6. Compute / Vwdr where u « log (x 2 -f y 2 ) and C is the circle x 2 -f y 2 « 1. 
Jc 


du 


-dx — 


- dy if x 2 > y 2 . 


8. Ifv 
through (0,0,0). 


1 t 

V - , compute / 
r Jo 


(*.*.*) J 

V - * dr over some simple path that does not pass 
(*o.vo.*g) r 


TRANSFORMATION THEOREMS 


6. Simply Connected Regular Regions. The validity of several im- 
portant theorems op the transformation of surface and volume integrals 
presented in the following sections hinges on the regularity and connec- 



TRANSFORMATION THEOREMS 


SEC. 6] 


383 


tivity of domains of definition of functions appearing in the integrals. 
A careful characterisation of such domains is extremely involved and is 
quite out of place in this book, but in order to aid the reader in under- 
standing the circumstances under which the theorems in question are valid 9 
we give a qualitative discussion. 

We shall say that a given region is connected if every two points of it can 
be joined by a smooth curve that lies entirely in the region. A region is 
simply connected provided that every simple closed curve 1 drawn in its 
interior can be shrunk to a point by continuous deformation without cross- 
ing the boundaries of the region. 

Thus, the interior of a square is simply connected, but the interior of a 
ring bounded by two concentric 
circles C x and C 2 is not (Fig. 15) be- 
cause a closed curve C surrounding 

cannot be shrunk to a point with- 
out crossing Also, the interior of 
a sphere is simply connected, and so 
is the interior of the region bounded 
by two concentric spheres, but the 
interior of a torus (an anchor ring) 
is not simply connected. A region 
that is not simply connected is called 
multiply connected . 

In dealing with bounded three- 
dimensional regions we shall say that 
the bounding surface S is smooth if 
at each point P of the surface one 
can erect a normal n(P) which changes continuously as P moves along the 
surface. A surface that can be subdivided by smooth curves into a finite 
number of pieces each of which is smooth is called sectionally smooth or 
pieceunsc smooth. The surface of a cube is an example of a piecewise 
smooth surface. 

The surfaces which we shall consider have two sides, although not all 
surfaces are two-sided. A one-sided surface can be formed, for example, 
by gluing the ends of a long strip in such a way that the upper side of one 
end of the strip is joined onto the under side of the other end (Fig. 16). 

If two oppositely directed normals PN and PN r are drawn at any point P 

of the surface, then the normal PN when carried along the path PABCP 

will coincide with PN'. It may be noted that this surface has a simple 
closed curve as its boundary. 



1 We recall that a simple closed curve is a closed curve consisting of a finite number of 
nonintersecting smooth curves. 


384 


VECTOR FIELD THEORY 


[CHAT. 5 

We shall suppose that all surfaces with which we deal are two-sided, 

piecewise smooth, and such that for 
some orientation of cartesian axes the 
projections on the coordinate planes 
consist of the interiors of simple closed 
curves. Such surfaces we shall term 
regular. If a region is a union of 
finitely many regions each bounded by 
a regular surface, it will also be called 
regular. 

Regions bounded by a cone, a 
sphere, or a cube are regular simply 
connected regions. The interior of a 
torus is an example of a regular 
multiply connected region, 

7. Divergence. Let a continuously differentiable vector point function 
v(P) be defined in a regular simply connected region R bounded by a 
closed surface <r. The surface integral of the component of v in the direc- 
tion of the exterior unit normal n(P) to a is called the flux of v over a. 
Thus, the flux F is 

F = £v*n d<r. (7-1) 

When v is the velocity of an incompressible fluid, the scalar F represents 
the amount of fluid issuing from a- per unit time. The points of the region 
at which the fluid is generated are termed sources , and those where it is 
absorbed are sinks. When the total strength of the sources is greater than 
that of the sinks, the flux is positive; when the strength is less, the flux is 
negative. 

Consider, now, a volume element r containing within it a point P } and 
denote the bounding surface of r by a. Then the flux of v over <r per unit 
volume is 

j vender 

(7-2) 

T 

If we let the volume r shrink to zero in such a way that the maximum 
diameter tends to zero, the quotient (7-2) will have a limit called the 
divergence of v at P. We denote the divergence of v by div v(P) so that 

|v-ni(r 

div v(P) « lim — 

t -> o r 

This quantity is a measure of the strength of the source at P. 



(7-3) 



TRANSFORMATION THEOREMS 


385 


EEC. 7J 

Inasmuch as the volume r is arbitrary, the existence and, indeed, the meaning of the 
limit in (7-3) are not quite obvious mathematically. One may let r approach zero while 
staying similar to itself, or one may let r become arbitrarily thin compared with its 
length, and so on. It is tolerably clear when suitable restrictions are imposed that 
all these processes yield a unique limit L, independent of the shape of r. Moreover, 
the convergence is uniform in the following sense: Given any « > 0, there is a 8 > 0 
such that 

^ v • n d<? 

L < «, (7 -&*) 

T 

provided the maximum diameter of r is less than 5. For rectangular solids r this fact 
is established in the next few paragraph, though the proof in the general case is not 
presented here. 

To calculate div v in cartesian coordinates we consider & volume r in 
the shape of a rectangular parallele- 
piped with center at P(x,y,z) and 
with edges Ax, Ay, A z (Fig. 17). 

The flux of v over the surface of this 
parallelepiped is easily computed. 

Since v = iv x + + kv z , the nor- 

mal component v*n of v over the 
face A BCD is v T . Hence the outflow 
over that face is (v x ) x+ i^^ x Ay Az, 
where (v x ) x +m Ax the mean value of 
v x over A BCD. Similarly, the out- 
flow over the parallel face EFGH is 

( — 0* )x-HAz Ay A z, 

where the minus sign appears be- 
cause the exterior normal to EFGH is — i and hence v*n = —v x . 

Thus the net outflow over a pair of faces parallel to the yz plane is 

*4 H At 

A y Az ss v x Ay A z. 

x — H Ax 



Proceeding in the same way with the remaining faces we get for the total 
outflow 

x+HAz v+MAy *4HAi 

I v*n do- = v x Ay Az -f v u Ax Az + v t Ax Ay. 

x — M Az y— HAj/ at — HAj 



z+H Az 
z—HAx 

Az 


div v(P) * lim 


V, 


(7-4) 




386 


VECTOR FIELD THEORY 


[CHAP. 5 


as Ax, Ay f and A z approach zero in any manner. Now, the three limits in 
(7*4) are the respective partial derivatives, so that we obtain the important 
formula 


div v(P) 


dv x 

dv,. dv g 

— + 

— H 

dx 

dy dz 


(7-5) 


The fact that the limits are partial derivatives is suggested by the definition of partial 
derivative (cf. Chap. 3, Sec. 2), Further discussion is required, however, because the 
functions v z , ty, v * are mean values. By the theorem of the mean [see Eq. (3-7), Chap. 3] 

G(x + MAx, y h zi) - G(x - M Ax, yi, zi) „ ^ 



Ax 


where Q\ stands for dG/dx and where $ is between x — H Ax and i + If Q\ 

is continuous, then 

I <?i(4,J/i,*i) - Gi(x,y,z) | 
is as snDall as we please provided only that 

It - *1 < 'A &x,\v - Vi\ Ay,|* - *i| <H 
with Ax, Ay, and Az sufficiently small Hence the mean value 


Ay Az 


/■* + K a* rV+M Av 

/ / Gi(^yi,z\) dyi dzi 

J z~H Ac Jy~ l A L 


a Ay 


is as close as we please to Gi(x,y,z), and the limit is therefore G\{x,y,z ). Applying this 
result to (7-4) with G — v x gives dv x /dx for the first limit, and the others follow by 
symmetry. 

The analysis shows that we may let Ax, Ay, and Az approach zero in any manner. 
For instance, if Ax 0 first, then Ay —* 0, and finally Az —* 0, the volume becomes 
a plane, a line, and finally a point. On the other hand if we set 

Ax ** ah, Ay » bh , Az =* ch 

where a, b, c are constant, and let h — > 0, then the volume stays similar to itself. Not 
only is the same limit obtained in all such cases, but the departure from that limit is 
seen to be uniformly small, provided only that 

max)|Ai|,|Avl,|A 2 | | 

is small. The remarks made in connection with (7-3o) are thus verified in this case. 


In terms of the differential operator 


d d d 

V = i — + j f-k — 

dx dy dz 

introduced in Sec. 3, we can consider a symbolic scalar product 


V*v 


/ d d d\ 

1 » 1 - + i + k r ) ' 0®* + K + kv *) 

\ dx dy dzf 

di)'x dt)y dv g 

dx dy dz 



SBC. 7] TRANSFORMATION THEOREMS 387 

On comparing this with (7-5) we see that 

divv=V«v. (7-6) 

We can also define the Laplacian operator V 2 by the formula 



and observe that if v = Vu y then 

div Vu = V-Vu = V 2 u. (7-8) 

Furthermore, if the symbol V x v is defined by the rule for computing 
vector products, we get 



It is worth observing that the condition V x v = 0 requires that each 
component of the vector V x v be zero. We can therefore write Eqs. 
(5-17) and (5-18) (which ensure the existence of a scalar u such that 
v — Vu) in the compact form V x v = 0. 

In Sec. 13 we shall deduce a formula for div v analogous to (7-5) when 
the vector v is referred to an arbitrary orthogonal curvilinear coordinate 
system. It is important to note that the definition (7-3) is independent 
of the choice of coordinates, so that div v is an invariant. 

Example ■ If v » i3x 2 4 j5 xy 2 4 k xyz*, compute div v at (1,2,3) and V x vat (x,y,z). 

Since v x *» 3x 2 , v y ** 5 xy 2 , v x «. xyz the substitution in (7-5) yields div v * 6x 4* 
10 xy 4* 3 xyz 2 . At the point (1,2,3) div v * 6 4 20 4 54 * 80. If v is interpreted as 
the velocity vector of fluid particles, we conclude that the point (1,2,3) is a source of 
the fluid. 

To compute V X v we use formula (7-9) and find 


V x v *» i(xz 3 — 0) 4 j(0 — yz z ) 4 k(5j/ 2 — 0). 


Since this vector is not identically zero, we conclude that no scalar function u(x,y,z) 
exists such that v » Vu. 



VECTOR FIELD THEORY 


[CHAF. 5 


PROBLEMS 

1 Find div v if (a) y « te -f )y -f kz t (b) v « i(x/r) + j(v/r) 4* kfc/r), where r « 

Vs* 4 v 2 -f aft**#, (c) v « i(* - y) -f j(s - g) ± k(y - x). 

%. Compute V 2 (I/r) and V 2 r, where r * Vx 2 -f ?/ 2 -f-z 2 . 

3. Shovfthat (a) div (u -f v) * div u -f div v, (6) div (uv) — V»(mv) * Vu*t + wV*v, 
(c) div (u K V) **■ V • x v) — v«(V x u) — u*(V x v). 

4. Show't&at div (r * a) « 0 if r « ix 4 iv *f k* and a is a constant vector. 

5. Find div (uv) if u *» x % 4" I/ 2 + s 2 and v — ir -f jy + k*. Also find div (Vu X v). 

8. The Divergence Theorem. An 

important relationship connecting 
the surface integral (7-1) for the flux 
of a vector field with the volume in- 
tegral of its divergence is deduced in 
this section. The resulting integral 
transformation theorem, known as 
the Gauss or divergence theorem, is 
fundamental to all developments in 
mechanics of continuous media. 

Let a continuously differentiable 
vector function v(P) be defined in 
a regular simply connected region r 
bounded by the surface <r. We sub- 
divide r into k volume cells Ar t in 
the shape pf rectangular boxes and parts of boxes (Fig. 18) and compute 
the divergence 

- i . . S^' n)da 

. , , div v(P,) = lira — (8-1) 

*’ ' Ar. ~+0 Ar t 

for each cel) v Ar»» (The role of r and a in (7-3) is now taken by Ar t and 
Affo.j Qn ^ajliiig the definition of limit, we can rewrite (8-1) in the form 

/ V-ndor « (div v) t Ar, + e.-Ar*, (8-2) 



where the h as A r x — > 
form the sum ’ ' t 

V ' 1 ‘ ' j: f v*n da 

. JA<r i 


0 and where (div v) t ss div v(P f ). 

k k 

= 2 (div v), An + a At, 

l=»l lw»l 


We next 


(8-3) 


over all the cells jpd observe that the surface integrals in (8-3) over the 
interfaces of adjacent cells vanish, since the exterior normals n to the 
common faces of ithe boxes point in opposite directions. Thus the surviv- 
ing terms in the sum on the left in (8-3) correspond to surface elements 



SEC. 8] TRANSFORMATION THEOREMS 889 

belonging to the exterior surface a, and hence this sura is equal to /v-n 

* . . r *" 

The sura 23 (div v), Ar» approximates the volume integral / div v dr, 

l-l * T 

and indeed, the approximation can be made as close as we wish by suitably 
decreasing the size (that is, the maximum diameter) of the ce$s. x T^e 

k 

sura of terras 23 & r % involves products of small quantities e* and Ar*, 

» — 1 

and it becomes arbitrarily small 2 in the course of the process described. 
We thus conclude from (8-3) that < . . 


J v*n da =* J div v dr. 


(8-4) 


The result embodied in this formula is the divergence theorem . This theott&ft 
expresses certain surface integrals as volume integrals, and since it con- 
tains no reference to any special coordinate system, the result is true in all 
coordinate systems. In particular, if v and n are expressed in terms of 
their cartesian components 


n = i cos (x,n) + j cos (y,n) + k cos (z,n), 
we can write (8-4), on recalling (7-5), as 

r r(dv T dv v dv z \ 

/ [v x cos (r,n) + v y cos (; y,n ) + v z cos (z,ri ) ] da = / ( i 1 ) dr. 

Ja J r\dx dy dz / 

(8-5) 

Example^ l. Verify the theorem (8-4) for v — i{xh) + j (y/r) -f k(z/r ), where r =* 
vV 2 -j- y l -f z 2 and the region r is the sphere x 2 + y 1 -t~ z l < a 2 . 

We readily find that 

dv x r 1 — J 2 dv y r 2 — y 2 dv z r 2 — z 2 ( f 

ox " '“T 2 ”’ Y y " ~ 7“' Yz “ “73“ ' 

80 that by (7-5) 3 ^ 2 __ ^ g 

div v = — « 

r r 


Now j div v dr » j - dr 1 * 

1 Of course, the number of cells k must increase without limit as this process is carried 
out. 

* This is true by virtue of the uniformity emphasized earlier [see (7-3a)]. Thus, given 
« > 0, we can make the subdivision so fine that | <»| < « for all the €,s aft bncbi 1 In that 
case ; ' * 

|2« t Ar*| < eS A r t *=> eV, t 


where V is the volume of the region. 


in* f 


I 



VECTOR FIELD THEORY 


390 


[chap. 5 


and it is easy to evaluate this integral in spherical coordinates, since in spherical coordi- 
nates dr m r 3 sin 6 do d4 dr (see Prob. 2, Sec. 2). We have 

f C 2 A* f r ^ 2 

/ div v dr ** I ~ dr «* 8 I I / - r 2 sin 9 d$ d*t> dr ■» 4*-a 2 , 

Jr Jrr Jo Jo Jo r 


On the other hand 


L v - adv ~L 


1 ‘da «• 47TO , 


since v*a « 1, for n « l(x/r) + j (y/r) + k (z/r) is directed along the radius of the sphere. 
Example 2. Prove with the aid of the divergence theorem the relation 


jvu dr — Jun d<r, 


(8-0) 


where u is a continuously differentiable scalar point function. 
Now in cartesian coordinates 

n * i cos ( x,n ) 4- j cos ( y,7i ) -f k cos ( z,n ) 

- i(n*i) + j(n-j) + k(n-k) 


and 


du da Ou 

Vu • i f~j f-k— , 

dx djy dz 


so that (8-6) is equivalent to the three equations 

f ~dr » [ (iu)-n d<r, 
Jr ax 4 

jT“dr » JiM-nda, 

lrTz dT ~ l ikU) ' adff - 


But these are the special cases of formula (8-5) applied to vectors v » iu, v — ju, v ** k u, 
and thus the correctness of (8-6) is established. 

Formula (8-6) can serve as a basis for a definition of Vu in the form 


Vu ■* lim 
r—* 0 


un da 
J<r 


(8-7) 


analogous to (7-3). 


PROBLEMS 

X. Prove that /r*n d<r * 3r, where r is the position vector of a point on the surface 
J<T 

of a regular simply connected region of volume r. Hint: Apply the divergence theorem 
to the surface integral. 

% Compute I vender, where o is the surface of the cylinder x 2 -f y 2 =■ o 2 bounded 
J<r 

by the planes * « 0, z « h, and where v » ix — jy + kz. 



SBC. 9] TRANSFORMATION THEOREMS 391 

3. Find j r*n dor, where r is the position vector of points on the surface of the ellipsoid 
(x 2 /a 2 ) 4* W/b 2 ) 4- (**. A* 2 ) - 1. 

4. Find the value of / v*n d<r, where v » r 2 (L r -f jy -f k*), r 2 « x 2 *f y 2 4~ z 2 , and 

J<r 

a is the surface of the sphere x 2 4- V 2 4* 2 2 * a 2 - Compute the integral directly and also 
with the aid of the divergence theorem. 

6. If v *** Vu and V 2 u ** p, where p is a specified scalar point function, show that 

it dff ^L PdT - 

Hint: Recall that 

du 

~~ « Vu*n. 
an 

i Use the divergence theorem to show that 

ir^ ( u Vv)dr ~ Ju Vv-nda. 

Show that this equation can be written as 

dv 


J uV 2 V dr Ju-j- dcr — J Vti-Vv dr. 


This important relation is known as Green's first identity. 

7. Using Prob. 6 obtain the symmetrical form of Green’s identity, namely, 


^(uV 2 !; — vV 2 u) dr * J (u -- — v d<r, 


which is also known as Green's second identity . (It is assumed in this identity that both 
u and v have continuous second derivatives.) Green’s identities are perhaps the most 
frequently encountered transformation formulas in mathematical physics. 

8. If the twice-differentiable function* u satisfies Laplace’s equation V 2 u «= 0, what is 

f du 

the value of / -~-d<r? Hint: Set v *» 1 in Green’s second identity, Prob. 7. 

Jar dn 


9. Green’s Theorem. Line Integral in the Plane. Because of the im- 
portance in applications of line in- 
tegrals defined over plane curves, 
we deduce here a special form of 
the divergence theorem commonly 
called Green’s theorem in the plane. 

Let a vector function 

v = iv x (x,y) + }v v (z,y) (9-1) 

with continuously differentiable 
components v X) v y be defined in the 
plane region R bounded by a simple Fra. 19 

closed curve C (Fig. 19). If we con- 
struct a right cylinder of height h with base R and apply formula (8-5) 




[CHAF. 5 


392 


VECTOR FIELD THEORY 


to the region r bounded by this cylinder, we get 


I [«?* cos (x,n) + v y cos (t/,n)] da 

*ir 



(9-2) 


The exterior unit normals n to the top and bottom bases of the cylinder 



thus write (9-2) as 


are k and — k, respectively, and 
hence cos (.r ,n) ~ cos (y f n) = 0 on 
the bases of the cylinder. The con- 
tribution to the surface integral in 
(9-2) from the bases is, therefore, 
zero, and the integral need be eval- 
uated only over the lateral surface. 
The element of surface da of the 
lateral surface is da — h ds , where 
ds is the arc element of C f and the 
volume element dr can be taken in 
the form dr = hdx dy. We can 


r f f / d v x dv y \ 

/ [r x cos (z,n) + v y cos (y,ri)]h ds= I 1 )hdy dx, (9-3) 

J c JJ R\dx dy/ 


where n is the exterior normal to C. But from Fig. 20 


cos ( x,n ) 


dy 

ds 


cos (y,n) 


dx 

ds 


(9-4) 


so that on dividing by h , Eq. (9-3) yields 


lc {Vl dy ~ V » dx) = Hr (5 + S) dX dy ’ 


(9-5) 


where in tracing C the region R remains on the left; that is, the path C 
is described in the positive direction. 

Formula (9-5) is Green's theorem in the plane. The function ~~v y (x,y) 
is sometimes denoted by M(x,y) and v x (x,y) by N(x,y), so that (9-5) 
assumes the form 

S c t Mdx + N dy) “ “ // fi (~ ~ ~) d * d y- (9*6) 

Our restrictions on v x and v y demand that M ( x,y ) and N(x,y) be continuous 
and have continuous partial derivatives in the plane region R. 

'We see that if dM/dy = dN/dx at all points of R, then J c (Mdx + N dy) 
0 over every simple closed curve C drawn in R. Conversely, if the 



SEC. 9] TRANSFORMATION THEOREMS 393 

line integral in (9-6) vanishes for every simple closed path C in R, then 


rr ( dM dN\ 

//«(■ 

for every region R. This enables us to prove that 

dM dN 
dy dx 

at every point of R; for suppose that 

dM dN 

5 ^ 0 


(9-7) 


( 9 - 8 ) 


(9-9) 


at some point P, and for definiteness let this difference be positive. Since 
(, dM/dy ) — (dN /dx) is continuous, there is a small region R' including P 
throughout which the integrand in (9-7) is positive. But this means that 
the integral is also positive, and since (9-7) is known to yield zero for every 
region R , we have a contradiction. Thus, the hypothesis (9-9) is untenable. 

We summarize these results as a theorem. 

Theorem. A necessary and sufficient condition for the line integral 
j c (M dx + N dy) to vanish for every simple dosed path drawn in a simply 

connected region R, where M f N, dM/dy, and dN/dx are continuous , is that 
dM/dy =*= dN/dx at all points of R. 

The vanishing of the integral 


f (M dx + N dy) (9-10) 

over every closed path is equivalent to the statement that this integral 
is independent of the path, and it follows from Sec 5 that the expression 
M dx + N dy is an exact or total differential du of a single-valued function 
u(x,y) determined by the formula [cf. Eq. (5-6)] 

u(x,y) = f ( >V) (M dx + N dy ). (9-11) 

J (zo,yo) 

We recognize condition (9-8) to be identical with (5-17). 

The theorem (9-6) can be extended to suitable plane multiply connected 
domains in the following way. If R is a doubly connected region bounded 
externally by a contour C 0 and internally by a contour Cj (Fig. 21), we 
introduce a “cut” C joining some point P 0 on C 0 with Pi on C\, The cut 
C can be visualized as a slit in the region R , forming the boundary C + C 0 
+ Ci of the slit region. The slit- region R is simply connected, and if we 



394 VECTOR FIELD THEORY 

apply formula (9-6) to it, we get 


[chap. 5 




* 4 ( Mdx + Ndy ) + f 1 (M dx + N dy) 

JC o JPo 

+ (£ (Mdx + Ndy) + f P °(M dx + Ndy ). (9-12) 
J Ci Jp i 



Fig 21 Fig 22 


The arrows on the integrals in (9-12) refer to the direction of integration 

rP , >0 

along Co and Ci as shown in Fig. 21, and the integrals I and / are 

evaluated along C in the direction indicated by the limits. Inasmuch as 
rP i rP 0 

Jp ~ ~~ J pi > Eq. (9-12) reduces to 


-u 


dM diY> 


= (M dx + N dy) + <j^ (ilf da* + A <&/). (9-13) 


An obvious extension of tliis result to the region 1? bounded externally 
by Co and internally by n contours C t (Fig. 22) yields 



4 (M dx + N dy) + X) 4 (M dx + N dy), (9-14) 
J Co ztaszl j a 

An important result follows directly from formula (9-14) if it is supposed 
that continuously differentiable functions M and N are such that 

dM dN 


dy dx 


(9-15) 



SEC. 9 ] TRANSFORMATION THEOREMS 395 

in the region R. If (9-15) holds in R f the double integral in (9-14) vanishes 
and we get 

<£ (Mdx + Ndy) « - £ <£ (Mdx + Ndy) 

J c '° ^ Ci 

« E<£ (Mdx + Ndy). 

»-i JCi 

Thus, the line integral over the exterior contour C 0 taken in the counter- 
clockwise direction is equal to the sura of the line integrals over the interior 
contours Ci taken in the same direction. In particular, if there is only 
one interior contour C \ (Fig. 21), we conclude that 

f (M dx + N dy) « [ (M dx + N dy). ( 9 - 16 ) 

•'Co JC\ 

This integral need not vanish. If, however, continuously differentiable 
functions M and N are also defined in the region interior to C\ and satisfy 
the condition (9-15) in that region, then the value of the integral 

J Ci (M dx + N dy) is zero, inasmuch as the integral on the left in (9-16) 
vanishes by theorem (9-6). 

PROBLEMS 


1. Show that the following integrals are independent of the path and find their values: 
fO.V 

(a) / f(x 2 *f y 2 ) dx -f 2 xy dy], 
o.i) 
rd.i) 


(6) r V f rr+% dx + (Tt* dy ] 1 

J (o.o) L(1 4~ x )* (J 4- x) 2 J 

Ar/2,r(2} 

(c) / ( y cos x dx 4~ sin x dy), 

J ( 0 , 0 ) 


id) 




X —1, 


X 2 < 1, 


r (2,3) 

(e) / (x -f 1) dx + (y + 1) dy. 

h i.i) r 

2. Write each of the integrals in Prob. 1 in the form / v»dr, and determine u(x,y) 


such that Vu ** v. 

3 . Find the value of 




— ydx 


+ 


x dy 


¥*)’ 


x 2 + y 2 ' x 2 + i 

where C bounds the region interior to the circle x 2 -f y 2 ** 4 and exterior to the circle 


x 2 4- v 2 « 1. What is the value of the integral (a) over the circle x 2 + y 2 
the circle x 2 + y 2 » l? 

4 . Compute the integral 


4? ( b ) Over 


!L 


div v dx dy, 


where v «■ ix + \y over the region R bounded by the circles x* 4* V 2 * 1 and x 2 4 - y 2 »■ 4, 



VECTOR FIELD THEORY 


[CHAP. 5 

5. tJ«e formula (9-13) to evaluate the integral I (~y dx x dy), where C is the path 

Jc 

bounding the region R in Prob. 4. What is the value of this integral over the path 
x l y % m 1? Over the path ar 2 -b y 2 «■ 4? 

10. Curl of a Vector Field. We saw in Sec. 7 that with every contin- 
uously differentiable vector function v(P) one can associate a scalar 
div v(P) defined by the formula 

j n«vd<r 

divv(P) * lim — (10-1) 

r — > 0 T 

which has a simple physical meaning. 

We show next that v(P) can also be associated with a vector field called 
curl v, defined by an analogous formula 

J n x v da 

curlv(P) = lim (10-2) 

r o T 

We shall see that curl v(P) bears an interesting relation to the concept 
of circulation in the vector field. 

Let v(P) be defined in some regular three-dimensional region P, and 
let C be a simple closed curve in R bounding a plane area A. At a given 
point P of A we construct a unit normal v so directed that v points in the 
direction of an advancing right-hand screw when C is traversed in the 
positive sense (Fig, 23). We then construct a right cylinder of small height 



h with elements parallel to v and with base A and denote its surface by cr 
and its volume by r. 



397 


SEC. 10 ] TRANSFORMATION THEOREMS 

Since v is a constant vector, formula (10-2) yields 

J v-n x v da 
vcurl v = lim — (10-3) 

r — » 0 T 

But along the bases of the cylinder v is parallel to the normal n, and hence 
the triple scalar product vn x v vanishes over the bases. Accordingly, 
the integral in (10-3) need be computed only over the lateral surface of 
the cylinder. We can thus write 

J vn x vhd8 

v • curl v ® lim — » (10-4) 

r o r 

since da « h ds. 

But vn x v * v*v x n by Chap. 4, Eq. (6-4), and v x n — t along C, 
where t is the unit tangent vector to C. Thus the integrand in (10-4) can 
be written 

vn x vhde — v*v x nh ds = v-t h ds = hv^dr 

where dr is the differential of the position vector r of a point on C. If 
we further note that r — hA , we can rewrite (10-4) as 

lc V ' dT 

v • curl v = lim (1U-5) 

a -+ o A 

The line integral is called the circulation of v along C. If v 

represents the velocity of a fluid, then v*dr = v*t ds takes account of the 
tangential component of velocity v and a fluid particle moving with this 
velocity circulates along C. A particle moving with velocity v-n normal 
to C, on the other hand, crosses C. That is, it flows either into or out of 
the region bounded by C ¥ . Hence formula (10-5) provides a measure of 
the circulation per unit area at the point P. This formula can be used to 
compute the cartesian components of the vector curl v by taking v suc- 
cessively as the i, j, k base vectors and by evaluating the limit in the 
right-hand member. It is somewhat simpler, however, to get the formula 
for curl v in cartesian coordinates from the definition (10-2) with the aid 
of the divergence theorem. 1 

Now the components of n x v in cartesian coordinates are n x vi, 
n x v*j, n x v*k. Consequently, 

n x v * i(n x v*i) + j(n x v*j) + k(n x vk) 

ss i(n*v xi)+ j(n-v x j) + k(n*v x k). (10-6) 

1 Also the uniformity of approach mentioned in connection with (7~3a) will then yield 
the same kind of uniformity for (10-2). 



398 VECTOR FIELD THEORY [CHAP. 5 

On inserting from (10-6) in (10-2) we get 

J (n*v x i) da j (n-v x j) do- 

curl v * i lim h j lim 

r — ► 0 r t ► 0 T 


J (n-v x k) da 

+ k lim — (10-7) 

r -* 0 T 

But a comparison of the right-hand member of (10-7) with (10-1) enables 
us to rewrite (10-7) in the form 

curl v =5 i div (v x i) + j div (v x j) + k div (v x k). (10-8) 


On inserting 


v = iv z 4- j v y + kv z 


in (10-8) we get 

curl v =» i div (jv e — kv y ) + j div (ky x — iv t ) + k div (iv v — jv x ), 


and a simple calculation making use of formula (7-5) yields the desired 
result 


/ dv z dVy\ ,/dv x dv z \ /dv v dv x \ 

cmU - ■ w ■ *) + ’ (* “ ta) + 1 (to - *)• <IM) 


If we recall the expression (3-13) for the symbolic vector V, we can write 
(10-9) compactly as 

i 3 k 


curl v = 


d d d 
dx dy dz 

V X Vy * V Z 


S5 V X V. 


(10-10) 


An analogous formula for curl v in orthogonal curvilinear coordinates 
is given in Sec. 13, and several useful relations involving the use of the 
curl operator are recorded in Prob. L 


Example: Compute curl v if v « ixyz -f j xyz 2 + kx?yz. 

The substitution of v x « xyz, v v — xyz 2 , v z » xhjz in (10-9) yields 


curl v — i(x®« — 2 xyz) -f j {xy — Zx 2 yz) -f k(yz 2 — xz). 



SBC. 10 ] 


TRANSFORMATION THEOREMS 


399 


PROBLEMS 


1. Show that under suitable hypothesis on continuity of the derivatives: 

(a) curl (A -f B) * curl A -f curl B; 

(b) div curl A «■ V»V X A *»* 0; 

(c) curl curl A« V div A — V 2 A, where V 2 A « i V 2 A Z -h j V 2 A v -f k 

(d) curl Vu ** V x (Vw) » 0; 

(e) curl ( uA ) - V X (wA) * uV X A -f Vu X A; 

(/) div (A X B) ** B*curl A— A-curlB; 

(g) curl (A x B) - AVB - BV-A + (B-V)A - (A-V)B, where (A-V)B m C is 
the vector with components 


dB x dB x dBx 

A*+A* + At ^ 
dx dy dz 



4 . Let a rigid body rotate with constant angular velocity ft about some axis through 
a point 0 in the body. If r is the position vector of a point P(x,y,z) relative to a set of 
axes fixed at 0, the velocity v of F is v ® v 0 + fi x r, where v 0 is the velocity of 0 rela- 
tive to some reference frame fixed in space (of. Sec. 8, Chap. 4). Show that curl v — 2ft, 
so that the angular velocity ft at any instant of time is equal to one-half the curl of the 
velocity field. Note that the velocity vo is independent of the coordinates (x,y, z) of 
points in the body. 

6. Show from geometrical considerations that the angle dd subtended at the origin by 
an element ds of a curve is do (n«r/r 2 ) ds. 

6. A solid angle w subtended by a surface o is measured by the area subtended by the 
angle on a unit sphere S with center at the vertex of <*>. Show that 


f 1 , 

(*> * — / n • V - d<r. 
Ja r 


where r is the position vector of points on <r measured from the vertex of u and n is the 
unit normal to a. Hint * Apply the divergence theorem to a volume formed by the bundle 
of rays issuing from the solid angle and by the areas cut out by these rays on S and on <r. 
7 . Referring to Prob. 6, show from geometrical considerations that 


do) ** — =- d<r. 

r 



VECTOR FIELD THEORY 


400 


[chap. 5 


tL Stokes’s Theorem. This useful integral transformation theorem 
enables one to reduce the evaluation of certain surface integrals to the 



calculation of line integrals. 

Let R be a three-dimensional re- 
gion in which v(P) is a continuously 
differentiable vector funotion and a 
a regular open surface embedded 
in R. We suppose that the edge 
of cr is a simple closed curve C 
(Fig. 24). Then it is true that 

J n • curlvder =*^v*dr, (11-1) 


where n is a unit normal to a and 
the line integral over C is evaluated in the direction determined by the 
chosen positive orientation of n. 

To establish formula (11-1), which is known as Stokes s theorem , we 
follow a procedure similar to that used in Sec. 8 to prove the divergence 
theorem. We subdivide a into k approximately planar elements of area 
A cr if each bounded by a simple contour C\ (say triangular) (see Fig. 24). 
Then formula (10-5) with v replaced by n % and A by A <r % when applied to 
the element bounded by C* yields 


curl v{Pi) A <T{ = / v«dr + e» A^. 

JCi 


(H-2) 


On summing these expressions over the entire surface cr we get 

k k k 

Y n i • curl v(P.) Act,- = Y / v-dr + Y «< A<r f . 

JGi l«*I 


(H-3) 


But the line integrals in (11-3) when summed over the common bound- 
aries of adjacent elements cancel out, since such boundaries are trav- 
ersed twice in opposite directions. The surviving terms yield the line in- 
tegral f c v*dr over the boundary C. If the number k of elements A <r t - 

is allowed to increase indefinitely, so that the greatest linear dimensions 
of the A<r* tend to zero, the sum on the left becomes the surface integral 

j n * curl v d<r. The sum J2 e, A or* tends to zero as in the discussion of 
J* i 

(8-3). 1 Thus, formula (11-1) is correct. It should be noted that once a 
positive direction for the normal n has been agreed upon, the positive 
direction of description of the contours C *•, and hence of C, is determined 
by the right-hand-screw convention. 

1 That is, if all | are lees than «, this sum is less than %S, where S is the area of a. 



SEC. 11] TRANSFORMATION THEOREMS 401 

If ir is a closed surface, the sum of the line integrals over the contours 
Ci is zero, and in that event jf n- curl v d<r « 0. 

We note further that if curl v « 0 in R, then L v*dr = 0 for an arbitrary 

rp J 

closed contour C. Hence the line integral / v*dr is independent of the 

path and thus defines a function u(P), such that du = v*dr. We can 
show conversely 1 that if the line integral in (11-1) vanishes for every 
closed path C in R, then curl v ~ 0 throughout R. Reference to (10-9) 
shows that the condition curl v = 0 is identical with Eqs. (5-17) and 
(5-18), ensuring that v*dr = du. 

Example: Evaluate / n ♦ curl v da over the surface t ** -f-Va 8 x 2 — y 2 if 

Jff 

v » i2y — jx -j- k*. 

The surface in this example is a hemisphere of radius a, and it is clear that 

. X V z 

n ■* l — j_j — f-k-! lz 

r r r v 

"" * y /* 

where r * ix + jt/ -f kz is the position x \ 'y 

vector for points on the hemisphere. We A 

readily check that 


dx dy dz 

I 2 y —x z I 

Hence f n • curl v da « — 3 / - da. 

Jcr J<j a 

This integral can be easily evaluated by 
noting that (Fig. 25) da « sec y dx dy, 
where y is the angle between the normal 
n and the positive direction of the z axis 
(cf. Chap. 3, Sec. 17). But from Fig. 25, 
sec 7 sec 0 *** aft, so that 



3 JJ dx dy 


since the region of integration A is a circle of radius a. The reader will check this result 
by taking da « a 2 sin 6 do d4> as the element of area of the surface of the sphere in 
spherical coordinates. 

To obtain the result (11-4) from Stokes's theorem (11-1) we compute / v«dr, where 

JC 

C is the boundary of the circle x z + y 2 * a 2 . Since dt » i dx + j dy -b k dz, we have 


/ v»dr * / ( 2ydx-~xdy+z dz). 
Jc Jc 


1 See the corresponding discussion in Sec, 9. 



402 VECTOR FIELD THEORY [CHAP. 5 

But along C we have di =* 0, and the equation of C may be taken in the form 


We thus get for (11-5) 



x ® a cos </>, 

y » a sin 4>> 0 <£ <f> < 2*. 

2 r 

a 2 ( 2 sin 2 <f> f cos 2 4>) d4> 



— 3ir a 2 . 


PROBLEMS 

1. Show that for the special case of a plane region bounded by a simple closed curve 
C, Stokes’s theorem reduces to Green's theorem (9-6). 

2. If v * iy -b jz -f kx and a is the surface of the paraboloid * « 1 — x 2 — y*, 

z > 0, compute / n * curl v d<r. 

f 

3. What is the value of the surface integral / n • curl v da if v « i y l -f- fay 4* kxar 

J<r 

and a is the hemisphere x 2 4* y 2 4- z 2 = 1, z > 0? Evaluate this integral directly and 
by Stokes’s theorem. 

4. Compute / v*dr if v « i(x 2 -f y l ) 4* ){x 2 -f z 2 ) 4- k y and C is the circle x L + y 2 

Jc 

4 in the plane z — 0. 

IS. Prove that the area A of the plane region bounded by a simple closed curve C in 
the xy plane is given by 

A “ j fj x dy ~~ y dx) 

when C is described in the positive direction. Hint: Use Green’s theorem (9-6). 

6. Verify Stokes’s theorem if v » iy 2 ~f- j xy — k xz and <r is the hemisphere z = 


x 2 - y\ 


ILLUSTRATIONS AND APPLICATIONS 


12. Solenoidal and Irrotational Fields. Let a continuously differentiable 
vector function v(P) be specified in a region R. If curl v = 0 at every 
point of R f we say that v(P) is an irrotational vector field. If v(P) is such 
that div v = 0, the field is said to be solenoidal . The importance of 
solenoidal and irrotational vectors in applications derives from the fact 
that every continuously differentiable vector function v(P) defined in a 
regular simply connected region R can be expressed as the sum of two 
vector functions, one of which is solenoidal and the other irrotational. 
We do not prove this fact here because it depends on demonstrating the 
existence of solutions of certain partial differential equations , 1 and it would 
carry us too far in the study of potential theory. Accordingly, we limit 
our discussion to proofs of two basic theorems concerned with solenoidal 
and irrotational vector fields. 

1 See Prob. 6. A discussion of the system of equations in question is contained in 
M. Mason and W. Weaver, “The Electromagnetic Field/’ pp. 352-365, University of 
Chicago Press, Chicago, 1932. 



ILLUSTRATIONS AND APPLICATIONS 


403 


SEC, 12] 

Theorem I, A necessary and sufficient condition that a continuously 
differentiable vector v(P) be irrotational in a simply connected regular region 
R is that v « Vu, where u is a single-valued scalar function with continuous 
second derivatives . 

We suppose, first, that v =* Vu; then 


curl v * curl Vu * 


i 

j 

k 

a 

a 

a 

dx 


dz 

du 

du 

du 

dx 

ty 

Tz 



as follows at once on expanding the determinant and noting the equality 
of the mixed partial derivatives of u(x,y,z). 

Conversely, if we suppose that curl v = 0 in R, then it follows from the 
concluding paragraph of Sec. 11 that du = v*dr and hence v = Vu. 

Theorem II. The continuously differentiable vector function v(P) is 
solenoidal m a region bounded by a regular surface if } and only if, it is equal 
to the curl of some vector w with continuous second derivatives. 

Let us suppose, first, that v — curl w. Then 


div v = div curl w as 0, 


as follows from a simple calculation 1 making use of formulas (7-5) and 
(10-9). Conversely, if div v = 0, we show that a vector w can be con- 
structed such that v = curl w. It suffices to show that the system of 
equations curl w ~ v, or 

dw z dw y 
dy dz 


dw x dw z 
dz dx 


Vy, 


( 12 - 1 ) 


dw v dw x 
dx dy 


has a solution for w x , 


w v , w t whenever 

dv x dv v dv z 
dx dy dz 


0 . 


(12-2) 


1 See Prob. 16 , Sec. 10. 



404 


VECTOR FIELD THEORY 


[CHAP. 5 

We show how to construct one such solution in rectangular domains. 
If we take te* «■ 0, then the second and third of Eqs. (12-1) require that 


dw z 




= -v y (x,y,z), = v,(x,y,z). (12-3) 

dx dx 

On integrating (12-3) with respect to x and treating y and z as constants, 
we get 

w,~ - [ v v (x,y,z) dx + <t>(y,z), 

(12-4) 

w v = I v,(x,y,z) dx + +(y, z), 

Jx 0 

where 4> and \p are arbitrary differentiable functions of y and z. If we 
insert these solutions in the first of Eqs. (12-1), we get 


But from (12-2) 
so that (12-5) yields 


. r^+^fV 

^o\dV dz/ 


d<l> dip 


(12-5) 


dv v dv z dv x 

dy dz dx 


v * */ 
Jx 


'x dv x d(j> d\p 

dx - 1 

dx dy dz 

d<f> d\p 

Vz(x,y,z) - v x (x 0 ,y,z) + — • 

dy dz 


( 12 - 6 ) 


( 12 - 7 ) 


This equation can be satisfied by taking f = 0 and 
4>(y,z) = f v x (x 0 ,y,z) dy. 

•'I/O 

Thus, one solution of the system (12-1) and (12-2) is 

w x = 0 , 

w v - T v t (x,y,z) dx, 

JXQ 

v>x "= - f v v (x,y,z) dx+ f v x (x 0 ,y,z) dy. 

JXQ J VQ 

The proof clearly indicates that w is not unique. Indeed, if we take w 
with components given by (12-7) and add to it Vu> where u is an arbitrary 
scalar function with continuous second derivatives, then 

curl (w + Vu) = curl w 

inasmuch as curl Vu se 0. 1 

1 Conversely, if curl Wi « v, then curl (wi — w) » 0 and wi — w *• Vu by Theorem 
I. Thus every solution wi is representable in the form w 4- Vu, where w is the par- 
ticular solution found in the text. 



ILLUSTRATIONS AND APPLICATIONS 


405 


SBC. 13] 

We remark in conclusion that whenever the divergence and curl of a 
vector function v are specified in the interior of a regular simply connected 
region and the normal component of v is known over the surface bounding 
the region, then there is just one vector function v satisfying these condi- 
tions. This uniqueness theorem is important in many applications. The 
reader may prove it by following suggestions given in Prob. 7 below. 

PROBLEMS 

1. Show that v ** i2xyz -f jx 2 z -f~ kx 2 y is irrotational, and find u(x } y,z) such that 
v «* Vu. 

2. Show that v » i(z — y) -}- j(x — z) + k (y — x) is solenoidal, and find w (x,y r z) 
such that v ■» curl w. 

1 Is v* i (y 2 4- 2 xz 2 — 1) + j2 xy + k2 x 2 z irrotational? If so, find u such that 
v * Vu. 

4. Is v » i(xh — 2 xyz) + K X V — 3 x 2 yz) -f k {yz 2 — xz) solenoidal? If so, find a w 
such that v ** curl w. 

6. Prove that v -* r n r, where r — ix 4~ -h kz, is irrotational. Ib it solenoidal? 

6. Let w «* u + v, where u is irrotational and v solenoidal in a given suitably re- 
stricted region R. Then there exists a vector q such that v ® curl q and a scalar 4> 
such that u *» V<£. Show that 4> and q satisfy the following partial differential equations: 

V 2 <#> * div w, V div q — V 2 q « curl w. 

7. If v is a continuously differentiable vector function defined in a regular simply 
connected region R bounded by the surface a and if 

curl v - f(x,y,z), div v - g(x,y,z) 

in R and v*n «* h{x,y,z) on cr, show that v is uniquely determined in R by these con- 
ditions. 

Outline of the Solution . Assume that there are two such vectors, v » Vi and v «. vj. 
With w =» Vi — V 2 , show that there is a u such that w ■* Vu, and deduce V 2 u *= 0. 

By applying the divergence theorem to the vector uVu, show that / (Vu) - (Vu) dr « 0. 

Jr 

Since (Vu)-(Vu) > 0, this integral can vanish only if Vu as 0. 

13. 'Gradient, Divergence, and Curl in Orthogonal Curvilinear Coordi- 
nates. In this section we record the expressions for the gradient, diver- 
gence, curl, and Laplacian in orthogonal curvilinear coordinates. These 
can be obtained from the definitions (7-3), (8-7), and (10-2) in a manner 
so similar to that used to obtain formulas valid in cartesian coordinates 
that we dispense with the details of calculations. 

As in Sec. 1, we suppose that a transformation 

Vi =* y%(x i,*a,sa), i = 1, 2, 3, 

wherein the variables yi are cartesian, defines a curvilinear coordinate 
system x. We suppose that the coordinates x, are orthogonal so that the 
quadratic differential form (2-13) has the structure 

(ds) 2 * gn{dxi) 2 + Qnidx*) 2 4- 



406 VECTOR FIELD THEORY [CHAP. 6 

We denote the unit base vectors along the coordinate lines by ©i, © 2 , ©a 
and represent a vector v(P) in the form 

V =« ©i^i + ©2^2 + ©3^3* (13-1) 

The volume element dr formed by the coordinate surfaces x, — const 

and x t + dx x ~ const (Fig. 26) has 
the shape of a rectangular parallel- 
epiped with edges 1 ds l = \/ y n dz u 
Hence the areas da XJ of its faces are 

da X2 * ^911922 dxi dx 2f 

daiz = Vgngzzdxi dx 3i (13-2) 

da 23 = ^ 22(733 dx 2 dx 3) 
and its volume dr is 



dr * Vg n g 22 g Z3 dx l dx 2 dx 3 . (13-3) 

To compute div v we calculate the flux da over the surface of the 

volume element dr and divide it by its volume (13-3). A calculation like 
that performed in Sec. 7 yields the result 


1 f d(vih 2 h 3 ) d(v 2 h t h 3 ) 

div v = 1 f 

h\h 2 h 3 L dxi dx 2 


d(vsh l h 2 y 
dx 3 . 


(13-4) 


where h t V^. 

A similar but slightly longer computation also yields the formula 

, 1 f d(h 3 v 3 ) d(h 2 v 2 ) 1 

curl v = ©x 

h 3 h 2 L dx 2 dx 3 J 


r <KMi) 

d(h 3 v 3 y 

1 

+ ©3 T~7~ 

d(h 2 v 2 ) 


L dx 3 

dxi - 

hih 2 

- dx x 

dz 2 J 


(13-5) 


which can be written more compactly as 


curl v = 


1 

hih 2 h 3 


k iei 

h 2 e 2 

h 3 e 3 

d 

d 

d 

dx } 

dx 2 

dx 3 

h i»i 

h 2 v 2 

h 3 v 3 


( 13 - 6 ) 


1 See Sec. 2. 



BBC. 13] ILLUSTRATIONS AND APPLICATIONS 407 

Finally, the formula for the gradient of a scalar u(x 1 ,x 2 ,xz) ) as follows 
from (8-7), is 1 


Vu 


©i du ^ ©2 du e 3 du 
hi dxi h 2 dx 2 h% Oxq 


(13-7) 


Inasmuch as div Vu = V 2 u, it is easy to check that the substitution 
of v = Vu in (13-4) yields 


V 2 u 


1 

hih 2 h$ 


d /h 2 h% du\ d /hih-j 
dxi \ hi dxi / dx 2 \ h 2 



(13-8) 

In cylindrical coordinates defined by the transformation 

x ~ r cos 0, 

y ~ r sin 0, 

2 = 2 , 

the metric coefficients are 2 

011 = 1 , 022 = r ~> 033 = 1 , 

so that hi = 1, h 2 = r, /? 3 = 1. 

Accordingly, formulas (13-4) and (13-8) yield 

1 <9 (rv r ) 1 dv$ dv z 

div v = ( 1 * 

r dr r <30 dz 



1 d 2 u 

+ ?h? + 


d 2 U 

a?’ 


where v = ri v r + 0i vo + kv z 

r b 0 b k being unit vectors in the direction of increasing r, 0, and z (Fig. 3), 
In spherical coordinates determined by 

x = p sin 0 cos <j> } 

y = p sin 0 sin <£, 

z = p cos 0, 

1 Henceforth we shall use the symbol Vu to mean grad u in curvilinear coordinates as 
well as in cartesian. 

* See See. 2. 



VECTOR FIELD THEORY 


408 


[char. 5 


hi « 1, h% ® p, A$ * p sin 0, as follows from Prob. 2 in Sec. 2. On making 
use of (13-4) and (13-8) we find that in spherical coordinates 


div v 


1 d(p 2 r p ) 1 d(sin0tty) 1 dv+ 


dp 


p sin B dB 


+ 


p sin 6 d<l> 


V 2 u 


l V dp/ i \ de/ 

: h : — i b * 


1 


d 2 u 


where 


dp p 2 sin 0 dB p z sin 34 B d<t> 4 

v — p x v p + diV 0 + 

and pi, 8j, are the unit vectors in the direction of increasing coordinate 
lines shown in Fig. 4. 

PROBLEMS 


1. Write out the expressions for Vu in spherical and cylindrical coordinates. 

2 . What is the form of V 2 in parabolic coordinates for which* 

(ds) 2 - (u 2 4 r?)[(dn) 2 + (dv) 2 ] 4 uVCd^) 2 ? 

8. The force F per unit charge due to a dipole of constant strength p is 
F » ri(2p cos 0/r 8 ) 4 0 L (p sin 0/r 8 ), 
where r, 9 are polar coordinates. Compute div F and curl F. 

14. Conservative Force Fields. In the concluding sections of this chap- 
ter we illustrate the use of vector analysis in the treatment of several 
problems drawn from mechanics, hydrodynamics, and the theory of heat 
flow in solids. 

When a particle of matter is displaced along a path C in a given field 
of force F, the work W expended in moving it is determined by the integral 

W^fv-dT. (14-1) 

The integral (14-1), in general, will have different values for different 
paths joining the same two points in the force field. If (14-1) is independent 
of the path, the field F is said to be conservative. 

We show next that the force field determined by Newton’s inverse-square 
law of attraction is conservative. 1 According to Newton’s law a particle 
of mass m located at a point P is acted on by a force F whose magnitude 
is proportional to m and inversely proportional to the square of the distance 
r from P to the center of attraction 0. Thus, 

km 

F- r„ (14-2) 

r* 

i A similar discussion applies to electrostatic force fields determined by Coulomb’s 
law, innee the mathematical structures of Newton’s and Coulomb’s laws are identical. 



ILLUSTRATIONS AND APPLICATIONS 


SBC. HI 


409 


where r* is the unit vector directed from 0 to P. The positive constant 
k is determined experimentally; it clearly depends on the choice of units 
of measure of F. Physically the law ( 14-2) represents the force of attraction 
of the mass m at P by a unit mass located at 0. 

If we rewrite (14-2) in the form 

km 

F=-— r, (1 «) 


where r = rr x , and insert it in the work integral (14-1), we get for the 
work done in displacing the particle from P x to P 2 along the path C, 

r km 

W = Jc ~r*dr. (14-4) 

But r*dr * ]^d(t*r) = r dr , so that we can write (14-4) as 



The integral (14-5) is clearly independent of the path joining P x and P 2, 
and if we denote r(P 2 ) hy r 2 and r(P x ) by r x (Fig. 27), we can write 



The function u(P) in (14-6) is con- 
tinuous at all points except when r = 0, and since div Vu = V 2 zi, we readily 
find that the gravitational potential (14-6) satisfies Laplace's equation 


V 2 u « 0 

except when r = 0. 

The gravitational potential at a point P due to a continuous distribu- 
tion of mass of density p is defined by the integral 

r kp dr 

u(P) * / * (14-8) 

J T T 

where r is the distance from the element of mass dm ~ p dr to the point P. 



410 VECTOR FIELD THEORY [CHAF. 5 

The force of attraction of the unit mass located at P by the body is 
determined by the formula F = Vu. 

The study of the properties of the scalar function u(P) defined by 
(14-8) is in the province of potential theory, and we shall encounter it once 
more in Chap. 6. 

Example: Let us compute the gravitational potential u(P) of a thin homogeneous 
spherical shell of radius a at a point P whose distance from the center of the shell is R 
(Fig. 28). 



Fig. 28 


The potential at P can be computed by summing potentials of the ring-shaped ele- 
ments of matter bounded by the cones with the semi vertical angles 0 and 0 -f- do. The 
area of the zone intercepted by these cones is 2wa sm 0 a do, so that 


u(P) - [ 
Jo 


'* kp2ra l sin 0 do 


where p is the mass per unit area of the shell. 
From the cosine law of trigonometry 


and we can write (14-9) as 


Vtf + H‘ L - 2 aft < 


u(P) - 2 rkpa 1 f —j~= 
Jo v or H 


R 2 — 2 aR cos 0 


if R > a, 


-[ V(R+a ) 2 


'(a - R ) s ), if R < a. 


If P is outside the shell, R > a, and we have the result 

Avkpa 2 kM 


(14-10) 


where M m Ara z p is the mass of the shell. 



SEC. 15J ILLUSTRATIONS AND APPLICATIONS 411 

When P is inside the shell, R < a, and we get 

u(F) * 4 rkpa, (14-11) 

a constant. 

The result (14-10) can be stated as a theorem. 

Theorem. The potential {and hence the force of attraction F ■* Vu) produced by a thin 
spherical shell at a poml exterior to the shell is the same as if the mass of the shell were con* 
centrated at its center. 

The potential due to a solid sphere of constant density p at a point outside the sphere 
can be deduced at once from (14-10) by supposing the sphere to consist of thin concen- 
tric shells We conclude that this potential has the same form as (14-10) with M re- 
placed by the mass of the sphere. Accordingly, the force of attraction produced by a 
solid homogeneous sphere on a unit mass at a point P outside the sphere has the mag- 
nitude kM/R 2 . This force is directed toward the center of the sphere 

From (14-1 1) we see that the force of attraction at a point inside the shell is zero. 

The integral (14-8) becomes improper if P is within the solid, for in that case, r =* 
V(x - £)* 4~ (y — v) 2 4- (2 — f) 2 becomes zero when the integration variables (£,i?,C) 
coincide with the coordinates (x t y,z) of P . However, the concepts of potential and gravi- 
tational attraction can be shown to have a meaning even when P is a point in the 
interior of a homogeneous solid. 1 


15. Steady Flow of Fluids. Let C be a curve in the xy plane over which 
a sheet of homogeneous fluid of 
depth 1 is flowing. The lines of flow 
of the fluid particles are indicated in 
Fig. 29 by curved arrows, and we 
suppose that the flow pat fern is 
identical in all planes parallel to 
the xy plane. A flow of this sort is 
called two-dimensional. 

The problem is to determine the 
amount of fluid that crosses C per 
unit time. We denote by v the ve- 
locity of the fluid particles at a point 
P on C and compute the volume dV Fig. 29 

of fluid crossing an element dr of C 

per unit time. Since the depth of the fluid is 1, this volume is equal to the 
volume of the parallelepiped 

dV = k-v x di, 

where k is the unit vector perpendicular to the xy plane. The volume V 
crossing C per unit time, therefore, is 

V = [ k*v x dr. 

h 

1 See in this connection Sec. 20, Chap. 1, and I. S. Sokolnikoff, “Tensor Analysis,” 
sec. 89, John Wiley & Sons, Inc., New York, 1951, where it is shown that the potential 
u satisfies Poisson's equation V 2 u * —4 xp. 




[chap, 5 


412 


VECTOR FIELD THEORY 


But by Chap. 4, Eg. (6-2), 


k*v x dt 


0 0 1 
V X Vy 0 

dx dy 0 


v x dy — Vy dx, 


since v * ii>* + ]v v and dt « i dx + j dy. 
Accordingly, 


V 


j (»Z dy - Vy dx). 


(15-1) 


If C is a closed curve and the fluid is incompressible, the net amount 
of fluid crossing C is zero, because as much fluid enters the region bounded 
by C as leaves it. Thus a steady flow of an incompressible fluid is char- 
acterized by the equation 

J c ( v x dy — v y dx) * 0, (15-2) 


where the integral is evaluated over any closed curve C not enclosing the 
points at which the fluid is generated or absorbed. But Eq. (15-2) implies 
that — Vy dx + v x dy is an exact differential d^(x 7 y) of the function 


Moreover, 1 


*(x>v) = 


r(X'V) 
J (xa,yo) 


(— Vydx + v x dy). 


d * 
dx 


v y) 


d * 


v*, 


and v x and v v satisfy the condition 

d( — Vy) dv x 

dy dx 


(15-3) 

(15-4) 


(15-5) 


throughout the region R in which (15-2) holds. Equation (15-5) is a con- 
sequence of (15-2); it states, in effect, that there is no fluid created or 
destroyed in the region R. For this reason it is called the equation of 
continuity . Since 

dv x dv y 

div v 1 > 

dx dy 

we can write (15-5) in vector form as 

div v * 0, (15-6) 

which is consistent with the meaning attached to the symbol div v in 
Sec. 7. 


1 See Sec. 5. 



ILLUSTRATIONS AND APPLICATIONS 


SRC. 15] 


413 


The function ^(x,y) defined by (15-3) is the stream function, and the 
tracks of the particles of fluid, or streamlines, are determined by the equa- 
tion ^(x,y) *» const. The velocity field satisfying (15-6), we recall, is 
said to be solenoidal. If the flow v is irrotational, then curl v **= 0 and 
there exists a scalar function $(x,y) such that 1 


(15-7) 
(15-8) 

The function $(x,y) determined by the integral 

r(x>v) rP 

$(x,y) « / (v x dx + v v dy)ss v-dr 

•'(*«.yo) JPo 


V = 

V4>, 


d$ 

— , 

Vy 

dx 

dy 


is called the velocity potential because of the relations (15-8). We emphasize 
the fact that the condition for the existence of $(x,y) is 


or in scalar form 


curl v = 0 
dv z dv v 
dy dx 


(15-9) 


If the flow is both irrotational and solenoidal, the relations (15-4) and 
(15-8) hold and we conclude that 


— « — • — = in R. (15-10) 

dx dy dy dx 

These are the celebrated Cauchy-Riemann equations which we shall en- 
counter again in Chap. 7. 

Furthermore, if div v = 0 and v is given by (15-7), we see that 

div V4> - V 2 4> = 0. (15-11) 

Thus, the velocity potential 4> satisfies Laplace's equation throughout any 
region containing no sources or sinks. 

On differentiating the first of Eqs. (15-10) with respect to y and the 
second with respect to x and on equating d 2 $/dx dy to d 2 $/dy dx , we find 
that the stream function ^( x,y ) also satisfies the equation 

V 2 * = 0. 

The practical importance of these results is stressed in Chap. 7, Secs. 19 
to 21. 

The foregoing considerations can be extended to the three-dimensional 
flows as indicated in Sec. 17. 


1 See Sec. 12. 



414 


VECTOR FIELD THEORY 


[CHAP. 5 


PROBLEMS 

1. Show that the gravitational field determined by (14-2) is both solenoidal and 
irrotational except at (0,0,0). 

2. Show that the velocity field 

. x __ y 

V x 2 -f y 2 J x 2 4 y 2 

is solenoidal in any region which does not contain the origin (0,0). Is it irrotational? 
Verify that the velocity potential <I> =» log r « log ( t 2 4 y l ) and the stream function 

m tan” 1 (y/x) ** 0. Compute the circulation around a circular path enclosing the 
origin, and thus obtain a physical interpretation of results m Probs. 5 and 6 of Sec. 5 
and Prob. 3 of Sec. 9. 

8. Discuss a two-dimensional flow for which the velocity potential <t> « cx . What is 
the stream function ^ for this flow? Plot the curves 4’ — const and & — const. 

4. Discuss a two-dimensional flow for which the stream function is 4' « 2 xy. Find 
the velocity potential <I>, and sketch the curves 4> - const and «= const. 

6. If v and w are irrotational vector fields, show that v x w is solenoidal. 

6. Show that the streamlines are orthogonal to the lines <f> » const. 

7. Show that when the three-dimensional flow v is irrotational, the stroamlines satisfy 
the equations 

dx dy dz 
v z v, 

8. If the velocity potential of the two-dimensional flow is 4> « x 2 — y 2 , find v and 
obtain the equations of the streamlines. Is this flow solenoidal? Is it irrotational? 

9. Show with the aid of the Cauchy-Riemann equations that when the stream func- 
tion ^(x ,y) is given, the velocity potential is determined by 


<*>(*, y) 


r(x, V ) v 

f ( — dx~-~-dy\- 
(*o.wo) dx ' 


19. Use the result given in the preceding problem to calculate «f >(x,y) if (a) Sk = x 2 — 
&zy 2 , (6) ♦ * — y/{x 2 4 y 2 ), x ^ 0, y ^ 0. 


16. Equation of Heat Flow. The following derivation of the Fourier 
equation of heat flow illustrates admirably the use of the divergence 
theorem in mathematical physics. 

It is known from empirical results that heat will flow from points at 
higher temperatures to those at lower temperatures. At any point the 
rate of decrease of temperature varies with the direction, and it is generally 
assumed that the amount of heat AH crossing an element of surface A<r 
in At sec is proportional to the greatest rate of decrease of the temperature 
u; that is, 

du 

AH == k A<r At — 
dn 


Define the vector q, representing the flow of heat, by the formula 

q “ ( 16 - 1 ) 



SBC. 16] ILLUSTRATIONS AND APPLICATIONS 415 

where fc is a constant of proportionality known as the thermal conductivity 
of a substance. [The units of fc are cal/ (cm-sec °C).] The negative sign 
is chosen in the definition because heat flows from points of higher tem- 
perature to those of lower, and the vector Vu is directed normally to the 
level surface u = const in the direction of increasing u. 

Then the total amount of heat H flowing out in A* sec from an arbitrary 
volume r bounded by a closed surface a is 


r du 
1 1 fc — da 

ri/n. 


I q*n da, 

J a 


since q*n - — kdu/dn by (16-1). 

On the other hand, the amount of heat lost by the body r can be cal- 
culated as follows: In order to increase the temperature of a volume ele- 
ment by A u° y one must supply an amount of heat that is proportional to 
the increase in temperature and to the mass of the volume element. Hence 

du 

A H *Ss c Aw p Ar « c — A t p A r, 
dt 

where c is the specific heat of the substance [cal/(g °C)] and p is its density. 
Therefore, the total loss of heat from the volume r in At sec is 


r du 

' / — cp dr. 
J r at 


Equating (16-2) and (16-3) gives 


[ q-nda = — / 

J a Jj 


Applying the divergence theorem to the left-hand member of (16-4) yields 


r r du 

/ div q dr = — / — cpdr , 

Jr at 


and since q = — fcVw, the foregoing equation assumes the form 


div ( — fcVw) + cp — dr 
dt] 


Now, if fc is a constant, 
and (16-5) becomes 


div (kVu) = kV 2 u 


-kV 2 u + cp — )dr sO. 


(16-6) 



416 


VECTOR FIELD THEORY 


[CHAR. 6 


Since this integral must vanish for an arbitrary volume r and the integrand 
is a continuous function, it follows that the integrand must be equal to 
zero, for if such were not the case, r could be so chosen as to be a region 
throughout which the integrand has constant sign. But if the integrand 
had one sign throughout this region, then the integral would have the same 
sign and would not vanish as required by (16-6). 

Therefore, 

—kV 2 u + cp — = 0 
dt 


or 

where 


— - h 2 V 2 u, (16-7) 

dt 

h 2 ^~ 

cp 


Equation (16-7) was developed by Fourier in 1822 and is of basic impor- 
tance in the study of heat conduction in solids. A similar equation occurs 
in the study of current flow in conductors and in problems dealing with 
diffusion in liquids and gases. 

It follows from (16-7) that a steady distribution of temperatures is 
characterized by the solution of Laplaces equation 

V 2 u - 0. 


It was assumed in this derivation that, the body is free from sources and 
sinks. If there are sources of heat continuously distributed within r, 
then it is necessary to add to the right-hand member of (16-3) the integral 

JJ{x,y,z,() dr, 

where J{x,y,z,t) is a function representing the strengths of the sources. 
The reader will show that in this case one is led to the equation 

— = h 2 V*u + — . 
dt cp 

provided that the thermal conductivity of the substance is constant. 
Thus the presence of sources leads to a nonhomogeneous partial differential 
equation. 

17. Equations of Hydrodynamics. Consider a region of space containing 
a fluid, and let v denote the velocity of a typical particle of the fluid. 
The amount Q of fluid crossing an arbitrary closed surface a drawn in 
the region can be calculated by determining the flow across a typical 
element A a of the surface <r. A particle of fluid is displaced in At sec 
through a distance v At, and since only the component of the vector v 



f 


imC. 17] ILLUSTRATIONS AND APPLICATIONS 417 


normal to the element Aa contributes to the flow across this element, the 
amount A Q of the fluid crossing A a is 

A Q & pv* n A <t At, 


where p is the density of the fluid (Fig. 30). 

The entire amount Q of fluid flowing out of the volume r, which is 
bounded by a, in At sec is 


Q « At f pv-ndo. 

Jo 

On the other hand, the quantity of the 
fluid originally contained in r will have 
diminished by the amount 




for the change in mass in At sec is nearly equal to (dp/dt) At At, and the 
negative sign is taken because p is a decreasing function of t. 

Equating these two expressions for Q gives 


/. 


pv*n da 



(17-1) 


and the application 
this equation yields 


or 


of the divergence theorem to the left-hand member of 



Since the integrand is continuous and the volume r is arbitrary, one can 
conclude that 

dp 

— + div (pv) * 0. (17-2) 


This is the basic equation of hydrodynamics, known as the equation of 
continuity. It merely expresses the law of conservation of matter. 

It has been assumed that there are no sources or sinks within the region 
occupied by the fluid. If matter is created at the rate kp(x,y,z,t), then the 
right-hand member of (17-1) should include a term that accounts for the 
increase of mass per second due to such sources, namely, 

J kpdr . 



418 VECTOR FIELD THEORY 

In this event the equation of continuity reads 


[chap. 5 


dp 

dt 


+ div (pv) 


fcp. 


The constant of proportionality k is sometimes called the growth factor. 

The density p(x,y,z,t) of the fluid at the location (x y y,z) of the fluid 
particle depends on t explicitly and on x,y f z implicitly, since the particle 
coordinates change with time as the particle is displaced. Thus, 


dp dp dp dx dp dy dp dz 
dt dt dx dt dy dt dz dt 


(17-3) 


In this equation, dp/dt means the rate of change of density as one moves 
with the fluid, whereas dp/dt is the rate of change of density at a fixed 
point. 

Upon noting that 


dx dy dz 

i hj h k — 

dt dt dt 


dp dp dp 

and Vp = i hj h k — * 

dx dy dz 

we can write the formula (17-3) as 


dp 

dt 


dp 

h v* Vp. 

dt 


Substituting from 


(17-2) in (17-4) gives 
dp 

— — — div (pv) + v-Vp. 
dt 


(17-4) 


(17-5) 


But div (pv) = v* Vp + p div v (see Prob. 36, Sec. 7), so that (17-5) 
becomes 


dp 

dt 


— p div v, 


1 dp 

or div v ~ (17-6) 

p dt 


It is clear from (17-6) that div v is equal to the relative rate of change of 
the density p at any point of the fluid. Therefore, if the fluid is incompres- 
sible, the velocity field is characterized by the equation 

div v « 0. (17-7) 

If the flow of fluid is irrotational, then curl v « 0, and one is assured 



SBC. 17] ILLUSTRATIONS AND APPLICATIONS 

that there exists a scalar function $ such that 


419 


v = V<£. 


Substituting this in (17-7) gives the differential equation to be satisfied 
by $, namely, 

d 2 <l> d 2 <t> d 2 <i> 


„ a^ 

v 2 4> m 1 h 

dx 2 dy 2 dz 2 


0 . 


(17-8) 


The function 3> is called the velocity potential. A similar result was obtained 
in Sec. 15 for the two-dimensional flow. 

If the fluid is ideal, that is, such that the force due to pressure on any 
surface element is always directed normally to that surface element, one 
can easily derive Euler’s equations of hydrodynamics. Denote the pressure 
at any point of the fluid by p; then the force acting on a surface element 
Aa is —pn Aa> and the resultant force acting on an arbitrary closed surface 
ar is 



The negative sign is chosen because the force due to pressure acts in the 
direction of the interior normal, whereas n denotes the unit exterior normal. 

Let the body force, per unit mass, acting on the masses contained 
within the region r be F ; then the resultant of the body forces is 

f F p dr. 

Hence, the resultant R of the body and surface forces is 


R = / Fp dr — / pn dcr 

Jr J a 

= j Fp dr - J Vp dr> 


(17-9) 


where the last step is obtained by making use of (8-6). 

From Newton’s law of motion, the resultant force is equal to 

r d 2 t 

R== h ^ dT > (17-10) 

dt 2 


where r ® ix + jy + kz is the position vector of the masses relative to 
the origin of cartesian coordinates. It follows from (17-9) and (17-10) that 

/ t (f p — Vp-pl^-O, 

and since the volume element is arbitrary and the integrand is continuous, 



420 


VECTOR FIELD THEORY 


[CHAP. 5 


ft 

' dt 2 


F p ~ Vp. 


( 17 - 11 ) 


This is the desired equation in vector form, and it is basic in hydro- and 
aerodynamical applications. 

In books on hydrodynamics, the cartesian components of the velocity 
vector dx/dt are usually denoted by u, v, and w, so that 

dx , dx dy dz 

— ® xu + \v + kw ® l h j h k — • 

dt dt dt dt 

Since n, v t and w are functions of the coordinates of the point (x y y } z) and 
of the time t, it follows that 

ih 
if 


i 


/du dudx dudy dudz ^ 
Kdt dx dt dy dt dz dtJ 


+ j 


(: 


dv dv dx dv dy dv dz\ 

“j 1 1 ] 

dt dx dt dy dt dz dt/ 


( dw dw dx dw dy dw dz 

1 1 1 — 

dt dx dt dy dt dz dt 


:)■ 


du du du 

dn 

i dp 

1 uH v + 

* — w = F x — 

— 

dt dx dy 

dz 

p dx 

dv dv dv 

dv 

l dp 

1 uH v + 


, 

dt dx dy 

dz v 

p dy 

dv) dw dw 

dw 

1 dp 

1 u H v + 

— w = F g — 

— 

dt dx dy 

dz 

p dz 


Substituting this expression in (1 7-11) and setting F = iF x + j F v + k F e 
ead to three scalar equations, which are associated with the name of Euler: 


( 17 - 12 ) 


It is possible to show with the aid of these equations (and by making 
»ome simplifying assumptions) that the propagation of sound is governed 
ipproximately by the wave equation 

d 2 S 2 ~ 
di? ~ a *' 

tn this equation, a is the velocity of sound and s is related to the density 
) of the medium by the formula 

$ = — — 1, 

Po 

adhere po is the density of the medium at rest. 



CHAPTER 6 


PARTIAL DIFFERENTIAL EQUATIONS 




The Vibrating String 

1. Arbitrary Functions: One-dimensional Waves 425 

2. Derivation of a Differential Equation 431 

3. Initial Conditions 435 

4. Characteristics 440 

5. Boundary Conditions 442 

6. Initial and Boundary Conditions 446 

7. Dam pal Oscillations 449 

8. Forced Oscillations and Resonance 451 

Solution by Series 

9. Heat Flow in One Dimension 455 

10. Other Boundary Conditions. Separation of Variables 459 

11. Heat Flow in a Solid 463 

12. The Dirichlet Problem 467 

13. Spherical Symmetry. Legendre Functions 471 

14. The Rectangular Membrane. Double Fourier Series 474 

15. The Circular Membrane. Bessel Functions 480 

Solution by Integrals 

16. The Fourier Transform 482 

17. Waves in a Half Plane 486 

18. The Convolution Theorem 488 

19. The Source Functions for Heat Flow 491 

20. A Singular Integral 493 

21. The Poisson Equation 495 

22. The Helmholtz Formula 499 

23. The Functions of Green and Neumann 501 


423 



424 


PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 


Elliptic, Parabolic, and Hyperbolic Equations 


24. Classification and Uniqueness 604 

25. Further Discussion of Uniqueness 607 

26. The Associated Difference Equations 610 

27. Further Discussion of Difference Equations 612 

28. An Example: Flow of Electricity in a Cable 614 

29. Characteristics and Canonical Form 617 

30. Characteristics and Discontinuities 619 





Equations containing partial derivatives arise in many branches of 
mathematical physics. Fluid flow, heat transfer, wave motion, electro- 
magnetic theory, elasticity, quantum mechanics, nuclear physics, and 
meteorology are but a few of the fields that involve a stud}^ of such equa- 
tions. In this chapter we give representative examples, indicating some 
of the more important methods of solution. In contrast to the theory of 
ordinary differential equations, it will be seen that now the general solution 
is seldom sought. The main problem, rather, is to find that particular 
solution which satisfies the determinative conditions (the so-called initial 
values and boundary values) of the specific problem in hand. 

THE VIBRATING STRING 

1. Arbitrary Functions : One-dimensional Waves. A partial differential 
equation of order n is an equation containing partial derivatives of order n 
but no higher derivatives. For example, each of the three equations 

d 2 u 2 d 2 u du 2 d 2 u d 2 u d 2 u d 2 u 
dt 2 Q dx 2 dt a dx 2 dx 2 + dy 2 dz 2 ° 

is a partial differential equation of order 2. In this chapter we shall often 
use the subscript notation for derivatives, so that the foregoing expressions 
can be written more briefly as 

Un *= a 2 U xx , lit ^ CX U>xx ) Mix *4“ Uyy Hh (1-1) 

A function u that satisfies a given partial differential equation is called 
a solution of the equation. For example, the function 

u = cos x cos at (1-2) 

is a solution of the first Eq. (1-1), because (1-2) gives 

u x ** — sin x cos at, u t « cos x(-~ a sin at), 

u xx ** — cos x cos at, u ti « cos x( — a 2 cos at) « a 2 u m . 

425 



426 


PARTIAL DIFFERENTIAL EQUATIONS 


[chap. 6 


The reader will recall that the general solution of an ordinary differential 
equation contains arbitrary constants; for example, the general solution 
of y ft + y » 0 is 

y — Cx sin x + c 2 cos z , 

which has the arbitrary constants c x and c 2 . We shall see that many im- 
portant partial differential equations have solutions which contain arbitrary 
functions and, conversely, the elimination of arbitrary functions from a 
given expression often leads to a partial differential equation. 

As an illustration of this fact let 


u * f(x + y), (1-3) 

where / is an arbitrary differentiable function. If the argument of / is 
denoted by s » x + y, then 

u ~ f(z + y) « f(s) 

and the chain rule 1 gives 

du df ds df 

-m 

dx ds dx ds 


Similarly, u y ~ f{s), and hence u satisfies 

tl x “ U y 

for any and all choices of the differentiable function f. 

Conversely, let u(x,y) be a solution of (1-4). If we set a » x -f* V, then 
u(x,y) » u(x, s — j) ^ U(x t s). 


The chain rule gives 


_ . dx 

U x l'x — + 

dx 


ds 

dx 


v x + u. 


(1-4) 


and, similarly, u v » Substituting into (1-4) we get 

r x + v 9 - r/„ 

which shows that U x ** 0. It follows that U is a function of a only, 

V - f(s) - fix + y\ 

and hence the same is true of u. Thus, (1-3) follows from (1-4). 

For an example containing two arbitrary functions, let 

U =* /i(r) + f 2 (s), fi an d/ 2 differentiable, (1-5) 

where r and $ are the independent variables. Then U r = /i(r), and hence 

f/ ra ** 0. (1-6) 

1 The reader may find it advisable to review Chap. 3, Sec. 4. 



SBC. 1] THE VIBRATING STRING 427 

(The reader can verify that also U MT = 0.) Conversely, from (1-6) we have 

r (tfr) - 0 

ds 

so that U r is independent of s: 

U r — h(r), a function of r only. (1-7) 

If we write /i(r) = j h(r) dr, then Eq. (1-7) yields 

« 0 , 

dr 

so that U — /i(r) = / 2 ($), a function of s only. Thus, E7 has the form 
(i-6). 

An important example of the elimination of arbitrary functions arises 
from the situation shown in Fig. 1. If ^ is time, it is seen that fi(x — at) 



represents a wave form which propagates in the positive x direction with 
velocity a and w r ith no change m shape, that is, with no dispersion. In a 
similar manner, f 2 (x + at) represents a wave form which propagates in 
the opposite direction with velocity a. The most general one -dimensional 
w r ave without dispersion is a superposition of two such, namely, 

u * fi(x — at) + f 2 (x + at). (1-8) 

Suppose, now, that u{x,t) is given by (1-8), with f\ and f 2 twice dif- 
ferentiable. If we set 


x — at — r, x + at 
then u ~ /i(r) + / 2 ($), and by the chain rule 
du 

u x = — 
dx 


*> 


(1-9) 


d(/i+/a)*r , d(f l +f 2 )ds 

+ 


dx 


dr dx ds 

The reader may verify similarly that 

= f\(x - at)(-a) + f 2 (x + at) {a). 


fi(x — at) + / 2 (x + at). 



{chap. 6 


428 PARTIAL DIFFERENTIAL EQUATIONS 

Differentiating again gives 

u xx «■ fi(x — at) + fl{x + at), 
u, t = f[{x — at) (—a) 2 +fi(x + at) (a) 2 , 
and hence « satisfies the partial differential equation 

u t t = a 2 u xx . (1-10) 

We show conversely that every solution u(x,l ) of (1-10) has the form (1-8) and thus 
represents the superposition of two waves propagating with velocity a. The substitu- 
tion (1-0) gives 

u{x,t) * U(r,s) 

so that, by using the chain rule as in the previous discussion, 
u x » U r 4 u t —aUr 4 aU 9 . 

Differentiating again yields 


Uxx ** Urr 4 2 U r » 4 " U stf 
Uu * aHlrr - 2a 2 Ur* 4 a 2 u* 9 . 

If we substitute these values into (1-10) we get (1-6). As we have already seen, tins 
ensures that U has the form (1-5), and hence u has the form (1-8). 

Equation (1-10) is satisfied by the most general one-dimensional wave 
motion with velocity a; and conversely, every solution of (1-10) represents 
such a motion. For this reason (1-10) is called the wave equation . To- 
gether with its analogues in two and three dimensions, (1-10) is an impor- 
tant aid in the study of many vibration phenomena. 

Example: Standing Waves. The motion given by 

fi(x — at) « A sin k(x — at), A, k const, (1-11) 

represents a sine wave of amplitude A and wavelength X = 2r/k, moving to the right 
with velocity a. The period T is the time required for the wave to progress a distance 
equal to one wavelength, so that X = aT or 



Similarly, a motion described by 

fc(x 4 -at) ** A sin k(x 4 at) (1-12) 

represents a sine wave, of the same amplitude and period, moving with velocity a to 
the left. The superposition of (1-11) and (1-12) gives 


which becomes 


u « A sin k(x — at) 4 A sin k(x 4 cU) 
u *» (2 A cos hat) sin kx 


(M3) 



429 


SEC. 1] THE VIBRATING STRING 

when we recall the trigonometric identities 

sin k(x =fc at) « sin kx cos kai dt cos kx gin hat. 

The expression (1-13) may be regarded as a sinusoid sin kx whose amplitude 2 A cos kal 
varies with the time ( in a simply harmonic manner. Several curves of (1-13) are sketched 
in Fig. 2 for various values of t. The points nr /k remain fixed throughout the motion 



and are called nodes. Although the result was obtained by superposing two traveling 
waves, the wave form (1-13) does not appear to travel either to the left or the right, 
and (1-13) is said to represent a standing wave 

The number / of oscillations or cydcs made by the wave per unit time is called the 
frequency. From the definition of the period T t it follows that / » l/T, 


PROBLEMS 

1. If u * f(y/x) with / differentiable, show that 

XU x -f yUy = 0 for X 7* 0. 

2. Show by direct differentiation that u ** sin kx sm kat satisfies the one-dimensional 
wave equation for every choice of the constant k , and express this function in the form 
( 1 - 8 ). 

3. (a) By computing u x% u Vt and Uxy obtain a second-order partial differential equation 
for u = /i(x)/ 2 (.v). (fi) Show that your result us equivalent to (log « 0, and explain, 

4 . For many functions the chain rule applies even when the argument is complex. 
Assuming this, show that 

u « /i(x -f iy) -f/ 2 (x - iy), i 2 » -1, 

satisfies Laplace's equation u xx -f Uyy » 0. 

5. Lot f(x -|- iy) ** u(x,y) 4- w{z,y), w'here u and v are real. Using the chain rule, 
show that u and v satisfy the Cauchy-Rieniann equations 

u x « v y , Uy » —v x . 

6. Show that u » f(ay — fix) satisfies 

aUx 4“ 0 

if / is differentiable and a, d are constant. 


430 


PARTIAL DIFFERENTIAL EQUATIONS 


(CHAP. 6 


D? 


The Linear Equation with Constant Coefficients 

7. The operators DJ and D* are defined by 

d n ^ d n 

dx n v dy n 

and we agree also, for example, that 

( otDz 4" 0D y )u ® aD x u -f* &D y u m ctUx 4* §tuy. 

(a) If nii are constant use the result of Prob. 6 to solve each of the equations 

(D x — rn\D v )u *0, (D x — m<iD v )u ** 0. 

( b ) Show that both solutions obtained in (a) satisfy 

(D x — miDy)(D x - m 2 D v )u - 0 (1-14) 

Hint: Since m, are constant, Eq. (1-14) may also be written 
(D x — m 2 D^(D z — miD y )u « 0. 

(e) Deduce that a solution of (1-14), containing two arbitrary functions, is 

u = Fi(y 4- mix) 4- /'Vy 4~ m 2 x), »n ^ m 2 , 

w * Fi(y 4- mix) 4- x F 2 (y + mix), m 1 * m 2 . 

tfin*: Since the equation is linear and homogeneous, the sum of the two solu- 
tions in ( b ) is again a solution. The result for mi * m 2 may be verified by 
direct substitution. 

[Similar results hold in general. The solution of 

(D x — mD v ) r u ** 0 

can be shown to be 

u ** Fi(y 4- mx) 4- xF 2 (y 4- mx) 4 - ... 4 - x r - l F r (y 4 - mi) 

and the solution for several such factors is obtained by addition (cf. Chap. 1, Sec. 21). 
The process gives the “general solution” in that the number of arbitrary functions 
equals the order of the equation.] 

8. The fourth-order equation 

dSl d 4 U d*u 

dx A dx 2 dy 2 dy 4 ^ 

occurs in the study of elastic plates. Show that the general solution is 

u - Fi(y - ix) 4~ xF 2 (y - ix) -f F z (y 4- ix) 4. xF^y 4- ix). 

Hint: The equation may be written 

(D\ 4- 2 DlD* 4 - Dt)u - 0 
so that the decomposition into linear factors gives 

(Dx 4- iDy)(D z 4- iDy)(D x - iDy}(Dx - iD v )u - 0 . 

Use the result of Prob. 7. 



431 


SEC. 2] THE VIBRATING STRING 

9. As in Probs. 7 and 8, solve: 

(a) u xx - a 2 tf w - 0; ( b ) u xx 4* « 2^; 

(0 Wxz 4* Uyv ** (d) Uxx + Uyy s* 2l4dey. 

10. Consider the equation 

U xx *f 4 Uxy 5 Uyy - /(X,J/). 

(o) By the method of Prob. 7 obtain a general solution when / ** 0. 

(b) By assuming u «* cy A , where c is a constant to be determined, obtain a par- 
ticular solution when / « y 2 . 

(c) Similarly, obtain a particular solution when / ** x. 

(d) By addition of the results (a), (b), (c) obtain the general solution when / *» 
3/ 2 + x. 

11. As in Prob. 10, obtain the general solution: (a) 2 z xx + ^ " 1 ; (b) z xx — a 2 ^ 

« X 2 ; (c) Z xx -f 3 Try 4* 2 Zyy “T-f J/. 

2. Derivation of a Differential Equation. Consider a flexible, elastic 
string stretched between two supports on the x axis (Fig. 3). To obtain 



a differential equation for the motion, let v (x f t) represent the vertical 
distance from the point x on the x axis to the string at time £. We shall 
apply Newton's law, 

(Mass) (acceleration of center of mass) =» force, (2-1) 

to the short piece of string between x and x + Ax. 

The mass of the short piece is 

Mass = p As 

where p is the mean density and As the length. The vertical component 
of acceleration for the short piece is 

Vertical acceleration — — - 
dt 2 


if u is the height of the center of mass above the x axis. To compute the 
vertical component of force we let T be the tension, and we introduce the 
angle B between the tension vector and the x axis. By Fig. 3 the vertical 
component is 


Vertical force due to tension « (T sin B) 


- (T sin $) . 





432 PARTIAL DIFFERENTIAL EQUATIONS (CHAP. 6 

If there is an additional vertical force F\(x,t) Ax due to other causes, 
substituting into (2-1) yields 

d^fl 

p As— — (T sin 0) - ( T sin $) + F x (x f t) Ax. 

dr x+Ax x 

Upon dividing by Ax and letting Ax —> 0 we get 
ds d 2 u d 

P 7~ ~ (F s* n 0) + Fx(x f t) (2-2) 

dx dt 2 dx 


if the required derivatives are continuous. 

To obtain a simpler equation, note that the definition of arc length 
yields 1 

ds r /du\ 2 i H 

- *+UJ 


and also 1 


dx 


~ 1 


sin $ * tan 0(1 + tan 2 0) H = u x (l + u 2 ) H ~ u x , 


if Wfc <$Cl. Moreover, if the displacement u is small, we can consider 
T « const. Substituting into (2-2) yields the approximate equation 


pu tt = Tu xx + F x (x,t). 

This in turn may be written 

u tt = a 2 u xx + F(x,t), (2-3) 

where a «= VT/p and F(x,t) — p~ l Fi(z 9 t). 

Equation (2-3) will be considered in the sequel under the assumption that 
p, and hence a, is constant. 

When the force function F(x,t) is zero, the vibrations of the string are 
termed free vibrations . By (2-3) the equation for free vibrations is 

u tt ** a 2 u xx (2-4) 


and hence the solution has the form (1-8). According to the discussion 
in Sec. 1, the motion can always be regarded as a superposition of two waves 
moving with velocity 


a = 



(2-5) 


1 The symbol ^ (read “is asymptotic to") means that the ratio of the two sides tends 
to I* A discussion of this useful notation is given in Chap. 1, Sec. 2. 

•The fact that the string is flexible means that the tension vector is tangent to the 
string, so that 

du 
dx* 


tan B « slope of curve 



THE VIBRATING STRING 


SRC* 2J 


433 


in opposite directions* Later we shall determine the precise form of these 
waves by considering the initial state of the string, that is, the state at 
t « 0, together with the conditions at the end points, 2 = 0 and 2 « l 
Inasmuch as the constant a in (2-4) involves only the ratio !F/p, two 
strings may behave similarly even if made of different materials* For 
example, a string with density 2 p under tension 2 T behaves like a string 
with density p and tension T , since both yield the same value for a. An 
equivalence of two different physical systems such as this is sometimes 
called a principle of similitude. 


The study of similitude belongs to an interesting branch of mathematical physics 
known as dimensional analysis . Although a general development 1 will not be given 
here, we shall describe the underlying idea as it applies to (2-5). 

Equation (2-5) relates three quantities a, T, and p which are expressed in different 
physical units. Iu the mks system * 


f meters*] 

f kilograms | 

j" kilogram-meters 1 

L second J 

P L meter J > 

L (second) 2 J 


where the square bracket is used to indicate that the measuring unit, rather than the 
value, is being described. The value of a for use in (2-5) is the number of such measur- 
ing units, that is, the number of meters per second, and similarly for p and T. 

If we decide to measure lengths in centimeters rather than in meters, then the value 
of a will be increased by a factor 100. In other words, 100a cm per sec is the same as 
a m per sec. Similarly T will be multiplied by 100, but p will be divided by 300, since 
the length unit for p in (2-6) occurs in the denominator. (Indeed, p kg per m is clem ly 
the same as 0.0 Ip kg per cm ) Hence when a string has a wave velocity a, density p, 
and tension T in the old system (2-6), then the same string has velocity, density, and 
tension 

100a, 1007- (2-7) 


in the new system. Substituting into (2-5) yields 

floor 

100 a -» - 

\ p/100 

which is consistent with (2-5), as it should be. One does not get a contradictory result 
by measuring all lengths in centimeters rather than in meters. 

When we change meters into centimeters, we divide the unit of length by 100. More 
generally, one might divide the unit by an arbitrary positive constant a. The new 
values of a, p, and T would be, respectively, 


oa, aT, (2-8) 

a 

[compare (2-7)]. Similar changes may be made in the units of mass or of time. Equa- 

1 The reader is referred to P. W. Bridgman, “Dimensional Analysis,” Yale University 
Press, New Haven, Conn., 1931, and S. Drobot, The Foundations of Dimensional 
Analysis, Studia Math., 14:84-99 (1954). 

»The mks units for T can be found from Newton’s law (2-1), since T is a force. 



434 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

tkm (2-6) remains seif-consistent under such changes, as the reader can verify* The 
question arises: Is (2-5) the only functional relationship 

a~f(p,T) (2-9) 

which is consistent under such changes? If so, then we would have a proof of the func- 
tional relation (2-5), assuming merely that there is a functional relation of some kind. 

To investigate tins possibility, suppose (2-9) holds where / is an unknown function 
and where a, T t and p stand for the numbers of their respective units of measurement in 
(2-6). If the unit of length is divided by a, then (2-8) gives 


«o 



upon substitution iuto (2-9). Since a is arbitrary we may choose « ® p to find 


pa »/(l,pT). (2-10) 

If we now divide the unit of mass by / 3 , the value of a is unchanged but p and T be- 
come ftp and pT, respectively [see (2-6)]. Substituting into (2-10) yields 

Ppa ~fi\,p 2 pT). 

Upon choosing p 2 «* (pT)~ l , we get 

(pT)~V - /(!,«, 


so that 



( 2 - 11 ) 


where c ** /(1,1) is a constant, independent of a, p, and T. 

Finally, if we divide the unit, of time by y, the new values of a, p, and 7' are given by 
(2-6) as a/y, p, and T/y 2 . Substituting into (2-11) gives 


a 

y 



which reduces to (2-11) again. Thus, no new information is obtained by changing the 
unit of time, and the constant c in (2-11) cannot be found by dimensional analysis. 
But we can determine c by considering the limiting case of small oscillations. The 
partial differential equation (2-4) is then valid, and (2-5) shows that c * 1. 


PROBLEMS 

1. The displacement of a certain string is 

w(s,0 - fi(x - at) +/ 2 (x -f at). 

What is the physical meaning of the condition u(0,t) m 0? If u(0,t) m 0, express fi in 
terms of / 2 , and thus deduce 

u(x,t) « f % (x 4- at) - / 2 ( - x + at). 

% (a) Fi nd/(x - at) when f(z) -(14 x 2 )~ l ; when /(a) *■ sin kx; wh mf(x) » e x . In 
each case compute also fix — at). Hint'. Substitute x — at for x in the expressions for 
/(x), (6) If u(x,t) * f(x - at) 4 -fix 4* at), find u(0,t), u(x, 0), and u({,l/o) for each f(x) 
In part (a) of this problem. 



SEC. 31 THE VIBRATING STRING 435 

3. In the derivation of (2-3) we observed that sin 0 ~ u Xf but we used this result in 
the form (sin 0) x ~ (u x ) x . (a) By differentiating the exact formula 

sin 6 =* u x { 1 + u?)~^ 

with respect to x , show that (sin 0) x ~ u xx is correct provided v xx is bounded. Also 
show that the error is of the order of u\ in this case. (6) By considering u 1 -f- x, 
v ■« 1 + 2x near x «* 0, show that the equation u ~ v does not always enable us to 
conclude u x ~ v x . (In other words: If two functions approximate each other the de- 
rivatives need not approximate each other and a separate investigation must be given,) 

4 . Show that the small longitudinal vibrations of a uniform long rod satisfy the dif- 
ferential equation 

A JfA 
dt* * P dx 2 ’ 


where u is the displacement of a point originally at a distance x from the end of the rod, 
E is the modulus of elasticity, and p is the density. Hint: From the definition of Young’s 
modulus E , the force on a cross-sectional area q at a distance x units from the end of 
the rod is Eq(du/dx) t since du/dx is the extension per unit length. On the other hand, 
the force on an element of the rod of length Ax is pq Ax d^d/dt 2 . 

5 . If the rod of Prob. 4 is made of steel for which E - 22 X 10 8 g per cm 2 and whose 
specific gravity is 7.8, show that the velocity of propagation of sound in steel is nearly 
5.3 X 10 5 cm per sec, which is about sixteen times as great as the velocity of sound in 
air. Note that in the cgs system E must be expressed in dynes per square centimeter. 

6. Show that the differential equation of the transverse vibrations of an elastic rod 
carrying a load of p(x) lb per unit length is 


El 


dx* 


p{x) - m 


d*y 
dt 2 ’ 


where E * modulus of elasticity 

/ = moment of inertia of cross-sectional area of roil about a horizontal transverse 
axis through center of gravity 
m ~ mass per unit, length 

Hint: For small deflections the bending moment M about a horizontal transverse axis 
at a distance x from the end of the rod is given by the Euler formula M ** El d 2 y/dx 2 , 
and the shearing load p(x) is given by d(*M/dx* — ;>(x) 

3. Initial Conditions. In the previous section the wave equation 

u tt *= a 2 u xx , a — const, (3-1) 

was derived for small displacements of a uniform flexible string. Ac- 
cording to Sec. 1 the general solution of (3-1) is 

u(x,t) = fi(x - at) + h(x + at) (3-2) 

where f\ and /2 are arbitrary twice-differentiable functions. 1 We shall 

1 Actually, (3-2) is meaningful whenever f\ and ft are well defined, and hence, condi- 
tions of differentiability are not emphasized in the sequel. A nondifferentiable function 
(such as the function shown in Fig. 4) is regarded as being a “solution” of (3-1) if it can 
be approximated, with arbitrary precision, by smooth solutions. See: I. G. Petrovsky, 
“Partial Differential Equations,” p. 05, Cambridge University Press, New York, 1954. 



PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 

now see that these functions can be determined from the initial conditions, 
that is, from the conditions at time t ~ 0. It is convenient to regard the 
string as infinite and the conditions as given for — « < a: < », The 
effect of the end points x = Q and x — l will be considered in Sec. 5. 



Fxa. 4. Ordinates on the resultant wave are obtained by forming one-half the sum 
of the oppositely moving waves shown by the dashed lines. 


Case I. Initial Impulse 0. Assume that the string is released from rest 
and that the initial shape is given by a known function /(x). (Such a sit- 
uation arises when the string is plucked, as in a harpsichord.) In symbols, 

u(x, 0) = f(x), u t (x,0) = 0, (3-3) 

where the second Eq. (3-3) expresses the fact that the vertical velocity 
du/dt is initially 0 for each point x of the string. By (3-2) we get 

u t (x,i) = —afiix — at) + af 2 (x + at) (3-4) 

upon using the chain rule as in Sec. 1. Since u t (x, 0) = 0, Eq. (3-4) gives 

/»(*)-/«(*) 

after dividing by a. It follows that / 2 (a:) = fi(x) + c, where c is constant. 



SBC. 3] THE VIBRATING STRING 437 

Using this equality with x replaced by x + at we see that (3-2) may be 
written 

u(x, t) « fi(x — at) +fi(x + at) + c . (3-5) 

This step is sometimes puzzling when encountered for the first time; namely from 

fc(x) « fi(x) 4* c 

how can we deduce fz(x 4- at) * /i(x 4- at) 4- c? The conclusion follows because the 
first equation holds for all values of x (and the conclusion would not follow otherwise). 
One cannot simply set x «* x 4- at, because that would lead to ai «* 0. But one can 
reason as follows: We have fz(x) «= fi(x) 4* c for all x. Hence fz(s) « fi(s) 4* c for all 
8 , and the choice s — $ 4 - at yields the desired result. 

So far we have used only the second initial condition (3-3). To ensure 
the first condition, u{x, 0) = f{x), we set t = 0 in (3-5) and equate the result 
to f(x), thus: 

fi(x)+fi(x) + c~f(x). 

It follows that /i(x) = Hf(x) — Yic } and substituting into (3-5) gives 
the final answer: 

u(x,t) « Yf(x — at) + Yf(x + at), (3-6) 

The displacement v(x,t) in (3-6) is the sum of two waves, each of the 
form }if(x) y which travel in opposite directions with the velocity a . In- 
itially (that is, for t =* 0) these waves coincide, but with the passage of 
time they diverge, the wave f(x — at) moving to the right and the other 
to the left. In particular, if the waves are of finite extent, then any given 
point of the string is at rest in the initial position after the passage of both 
waves. The situation is illustrated schematically in Fig. 4 when f(x) is a 
triangular wave on (—k,k). 

Casb; II. Initial Displacement 0. Suppose, next, that the initial dis- 
placement is 0 but that the initial velocity is not 0. (Such a situation 
arises when the string is struck, as in a piano.) If the initial velocity is 
g(x) at point x of the string, the initial conditions are now 

u(x,Q) * 0, u t (x f 0) « g(x). (3-7) 

The first Eq. (3-7) gives 

fi(z) +/ 2 0r) = 0 , 

when we recall (3-2), so that /ate) = —fi(x) for all values of x. Using 
this equality with x replaced by x + at y we see that (3-2) may be written 

u(x y t) * f x (z - at) - fi(x + at). (3-8) 

Differentiating (3-8) with respect to t and setting t » 0 yield 

u,(x,0) = -afi(,x) - of i(x) - g(x) 



438 PARTIAL DIFFERENTIA L EQUATIONS [CHAP. 6 

when we use the second condition (3-7). It follows that 

fi(x) ~ (*g(8)ds + c, ( 3 - 8 ) 

2a 

where c is constant, and hence (3-8) gives the final answer 

1 rx—at 1 rx+at 

u{x,t) =-— g(s) ds + ~ 0(8) ds, 

2 a J o 2a J o 

The result may be expressed more compactly as 

1 rx + at 

u(x,t) = — / g(s) ds. (3-10) 

2 a 

Equation (3-8), like (3-0), represents a superposition of two waves 
traveling in opposite directions. Here, however, the shapes of the waves 

are determined by fi(x) and —fx(r), 
which are mirror images of each 
other in the x axis. Moreover, the 
shapes are not found directly by the 
initial condition but are obtained 
through the integration (3-9). For 
this reason the waves may be of infi- 
nite extent even when the initial 
impulse u t {x, 0) = g(x) is confined to 
a finite portion —k < x < k of the 
string. Indeed, for such a choice of 
g(x) formula (3-10) shows that any 
given point x of the string eventually 
suffers a permanent displacement 

“ l\g(s)ds . (3-11) 

2a J ~ k 

This is the case because when 
at > k + 1 x | , the interval (x — at, 
x -f at) contains the interval (—&,/:). 
Inasmuch as g(x) = 0 outside the 
interval (—&,&), the integral (3-10) 
is then equal to (3-11). Since each 
given point of the string eventually 
moves the same distance (3-11), the 
part of the string that is again at rest forms a straight line parallel to the 
Original string. It is most interesting that this happens regardless of the 
choice of g(x), provided only that g{x) « 0 outside some finite interval. 
Graphical illustration is given in Fig. 5 for the case g(x) =* I on (—k,k). 


i 

~k 

k 

FTTT 

[fTn 

* 

— 1 

^ 




Fig. 5 



THE VIBRATING STRING 


439 


SEC. 3] 

Case III. Arbitrary Initial C<fnditions. Suppose, now, that both the 
initial displacement and the initial velocity are given by arbitrary functions 
of Xf so that the initial conditions are 

u(xfl) = /(x), u t (xfl) * g(x). (3-12) 

This problem can be solved by superposition of the two solutions previously 
obtained. Indeed, let v(x } t) and w(x y t) satisfy the wave equation (3-1) 
and the respective initial conditions 

*(»,0) = f(x), v t (x, 0) = 0, 

w(x,0) « 0, WifoO) = g(x). 

Then the function 

u(x y t) = v(x,t) + w(x,t) 

satisfies the wave equation because v and w do, and addition of the relations 
(3-13) shows that v satisfies (3-12). Since the wave equation was solved 
in the previous discussion subject to initial conditions of the type (3-13), 
addition of the two solutions obtained formerly gives the solution desired 
now. That is, 

1 l 1 rz+at 

u(x,t) = -f(x - at) + -/(x + at) + — / g(s) ds . (3-15) 

2 2 2 a 

The expression (3-15) is known as d'Alembert's formula; it satisfies (3-1) 
and (3-12), hence gives the motion of a string subjected to arbitrary initial 
displacements and velocities. 

The formula (3-15) can also be used to find the displacement of a semi- 
infinite string (0 < x < <») fixed at x = 0. If the initial displacement and 
velocity of a semi-infinite string are 

w(x,0) « /(x), wt(x,0) = g(x), x > 0, (3-16) 

we can imagine an infinite string for which the initial conditions in the 
interval (0,«>) coincide with (3-16) and in the interval ( — «,()) are deter- 
mined by 

u(x,0) * -/(|x|), u*(x, 0) = -0(|x|), x < 0. (3-17) 

The point x = 0 of an infinite string, moving in accord with (3-16) and 
(3-17), will obviously be at rest, and the behavior of the infinite string for 
x > 0 will be identical with that of the semi-infinite string. 

The superposition method does yield a solution of the problem but does not establish 
the uniqueness of that solution* We shall now show that every solution of (3-1) and 
(3-12) can bo represented in the form t; + w t with v and w as in (3-13). Since v and w 
were already shown to be unique, it will follow that u is also unique. 


(3-13) 

(3-14) 



440 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

Indeed, let u(x,t) be a solution of the wave equation (3*1) which Satisfies the initial 
conditions (3-12). Let v(x,t ) be the unique solution of (3-1) satisfying the first condi- 
tions (3-13). Then the function w(x } t) defined by 

w{x t l) a* U(x,l) — v(z,t) 

satisfies (3-1) and the second set of initial conditions (3-13), as the reader can verify. 
It follows that w is uniquely determined and hence u(x,t) is also uniquely determined. 
Because of uniqueness, (3-15) describes the behavior of the string Without uniqueness, 
we could only say that (3-15) describes a possible behavior of the string* 

PROBLEMS 

1. The displacement of a string is given by the traveling wave 

u(x t t) ** sin (x — at). 

What are the initial displacement and velocity? Verify, by actual substitution into 
(3-15), that your initial values yield the correct result, u(x,t) » sin (x — at). 

%, For a freely vibrating string the initial displacement and velocity are, respectively, 
sin x and cos 2x. Find the displacement and velocity of the point x ** 0 when l ■* 
Nmt: First find u( x t f) from (3-15). 

3. A freely vibrating string was subjected to an initial displacement 6 cos 5x and 
initial velocity 0. One second later it is found that the point x ■» 0 is displaced three 
units from the equilibrium position; that is, u(0,l) = 3. What can you say about the 
velocity of propagation for waves on this string? 

4. The initial velocity of a freely vibrating string is arc""* 2 . For what choice of the 
initial displacement (if any) does the resulting motion represent a traveling wave travel- 
ing in the positive x direction? Hint: It is desired that u(x,t) «* f\(x — at). Determine 
/i from the initial velocity, and then determine the initial displacement from j\. 

6. Solve Prob. 4 with the words “velocity'* and “displacement" interchanged. 

6. A stretched infinite string is struck so that its segment —1 < x < 1 is given an 
initial velocity 1. Use (3-15) to find the displacement and sketch the displacement 
curves for t ** 1/a and i *= 2/a. 

7. The initial displacement and velocity of a semi-infinite string are u(zfl) sin x, 
u t (x f 0) » 0, 0 < x < ». Find u(: c,t) for t > 0. Also find u{x t t) if u(x, 0) — 0, u*(x,0) =* 
—2a cos x, 0 < x < oo. 

4. Characteristics. A physical interpretation may be given not only by 
plotting u(x,t) versus x for a succession of values of t, but also by consider- 
ing the xt plane. Each point of the xt plane represents a definite position 
on the string at a definite time t. If we take £ = 0 to be the present time, 
then the half planevS t < 0 and t > 0 give the past 1 and future, respectively. 

Since the speed of propagation is a, the disturbance at (x,t) will reach 
a point (x 0 ,*o) given by 

1 Although it is not appropriate to permit t < 0 when the string is plucked or struck 
at t ** 0, it is appropriate if the string has been in motion for some time and the initial 
conditions are determined by high-speed photography. We could then take the view- 
point that we are trying to ascertain the past history of the string by observations on 
the present. 



SBC. 4] 


THE VIBRATING STRING 


441 


X — Xq X — 

— a or « —a, 

t- to t- to 


( 4 - 1 ) 


for the direct wave f\(x — at) and the opposite wave f 2 (x + at), respec- 
tively. Equations (4-1) may be written 


x ~ at * Xq — af 0 , x + ~ x 0 + afc- (4-2) 

If we draw the two lines (4-2) through the point (xofy), as shown in Fig. 6 # 
their intersection with the x axis 
(that is, t - 0) gives those points on 
the string for which the initial con- 
dition contributes to the disturb- 
ance at (xo,k)). The lines (4-2) are 
called the characteristics of the par- 
tial differential equation (3-1). 

Along the first line (4-2), x — at 
is constant, and hence f } (x — at) is 
constant. Thus, the deflection due 
to the direct wave is the same at all points of the first characteristic (4-2). 
The second line serves the same purpose for the opposite wave, and we 
can say, briefly, that the disturbance travels along the characteristics. 

If the initial disturbance is confined to some interval (x lf x 2 ), then we 
have the situation shown in Fig. 7. The xt plane is divided by the charac- 




teristics into six regions. In region 1 the points receive the disturbance 
from both waves, in II only from the opposite wave, and in III only from 
the direct wave. The points in IV and V are too far away to receive any 
disturbance at the corresponding times, and the points in VI are at rest 
because both waves have passed. That is, if P is a point in the region 
VI, then the characteristics through P (shown dashed in the figure) in- 
tersect the x axis outside the interval (xi,x 2 ). Hence the initial displace- 
ment at these points is zero, and we need consider the initial impulse only. 



442 


PARTIAL DIFFERENTIAL EQUATIONS 



(CHAP. 6 

Since the characteristics intersect outside the interval (ri,^), the dis- 
placement at P due to the initial impulse is given by the constant value 
(3-11). 

We have seen that the initial conditions determine both the direct wave 
and the opposite wave at each point on the x axis where these condi lions 
are given. Since the disturbance propagates along the characteristics 
the following theorem is suggested : 

Theorem I. Let u and u t be given on the interval (ti,x 2 ) in Fig . 8, and 

suppose Utt = a 2 u XJ . Then u(x,l) is 
uniquely determined in the shaded 
region but is ?iot uniquely determined 
at any other point. 

Both the initial displacement and 
the initial velocity have to be speci- 
fied in Theorem I, just as one would 
expect intuitively. It is a remark- 
able fact that the displacement alone 
(without the velocity) will deter- 
mine the solution, provided this dis- 
placement is given along tw f o intersecting characteristics in the xt plane. 
Indeed, let u(x,t) be given along (x u P) in Fig 8. Since the direct w r ave 
fi(x — at) is constant on (x\,P) 9 we can ascertain the shape of the reverse 
wave f 2 (x + at) along (jti,P). This, in turn, gives f 2 (x + at) along (x u x 2 ) t 
because the disturbance f 2 (x + at) propagates from (x lr r 2 ) to (xj,P) along 
the characteristics parallel to (x 2 ,P) (see the dashed line in the figure). 
In just the same way, w r hen u(r,t) is given on (x 2f P), we can determine 
the shape of the direct wave f\ (x — at) on (xi,x 2 )- Thus, we are led to the 
following theorem: 

Theorem II. Let u be specified along the two intersecting characteristics 
(xi,P) and (x 2l P) in Fig . 8, and suppose that u fi = d z u xx . Then u(x,t) is 
uniquely determined in the shaded region but is not uniquely determined at 
any other point. 

Theorems I and II are the fundamental existence and uniqueness theorem 
for the wave equation, deduced here by physical considerations. A simple 
mathematical proof of the same results is given in See. 25. 


Fig. 8 


6* Boundary Conditions. We now suppose that the freely vibrating 
string is not infinite but is stretched between two points of support (Fig. 
3). When the supports are on the x axis and do not move, the situation 
is described by 

u(0,t) « 0, u(l,t) * 0 for all t (5-1) 

These are called boundary conditions , because they refer to the boundary 
points of the interval (0,1) in which our physical problem is defined. Al- 

rtKiriAnolv Hr* rmf mnfmn 



THE VIBRATING STRING 


443 


SEC. 5] 


uniquely, they do enable us to establish some of the most interesting and 
important properties of the motion. Hence, in this section we see what 
can be deduced from (5-1) alone. In the next section we use (5-1) together 


with appropriate initial conditions. 

Physically, one would expect the 
until the disturbance created by the 
ends reaches the point of observa- 
tion. In terms of Fig. 7, the ends 
x = 0 and x = l have no effect in 
the region I provided the points 
x = 0 and x = / lie outside the 
interval (xj^r 2 )- When the disturb- 
ance reaches an end point, however, 
it is reflected, and the reflected 
wave must eventually be taken into 
account. 

Because the end point is fixed the 
incident and reflected waves have 
algebraic sum 0 at the end point, 
and hence there is a 180° phase shift. 
A wave of type / 2 (.r + at) becomes 
a w^ave of type —/ 2 ( — .r + at) upon 
reflection at x = 0, for example (see 
Fig. 9). The change of sign m / 2 
expresses the phase shift, and the 
change of sign in x indicates that 
the reflected wave 

g(x - at) s -f 2 (~x + at) 


string to act like an infinite string 



Fig U 


propagates in the opposite direction. 

When the wave is reflected again at x — l, we get another minus sign 
in each case, and hence the original wave f 2 (x + at) is restored (Fig. 9). 
Since the velocity is a and the length of the round-trip path is 2 1, the time 
for a round trip is 

21 

Period of vibration — — • (5-2) 

a 


In terms of/ 2 (af) the periodicity condition means that 

/ 2 (tt/) = / 2 [a j = f 2 (at + 21 ). 

Similar remarks apply to /i(x), and hence we expect that both /i(x) and 
/ 2 (x) will be periodic functions 1 with period 21 

1 A function f(x) has period p if f(x -f p) m f(x), where p is a nonzero constant. 



444 


PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 

To discuss the boundary conditions mathematically, let us think of 
the finite string as being in reality an infinite string which vibrates in such 
a way that the points x « 0 and x » l remain fixed. The formula 

u(x,t) ® f\{x — at) + f 2 (x + at) (5-3) 

holds for all solutions of the wave equation. Letting x =* 0 gives 

0 * fi(-at) +/a(o0 

when we use the first boundary condition u(0,t) = 0. This shows that 
f 2 (s) « — /i( — s) for all s, and hence (5-3) becomes 

u(x,t) = fi(x — at) - at). (5-4) 

Thus the effect of the boundary condition at x = 0 is to reduce the number 
of arbitrary functions from two to one. 

The second boundary condition applied to (5-4) gives 

0 — /i(-i — a€) 

or, if we set s = —l — at, 

o «/t(« + 20 -/i«. (5-5) 

Since t is arbitrary, so is s, and hence f\ (x) has period 21. (This agrees 
with the surmise we had formed on physical grounds.) In view of (5-4), 
we can summarize our result as follows: 

Theorem I. Suppose an infinite string vibrates freely in such a way that 
the points x * 0 and x ~ l remain fixed . Then the displacement u{x,i) is 
periodic both in space and in time. The two periods are , respectively , 21 
for x and 21 /a for t if a is the velocity of propagation. 

Hence if a string is stretched between two fixed points, the free vibrations 
are periodic no matter what the initial conditions may be. Since a periodic 
vibration is generally perceived as musical, this fact is of great importance 
for the development of musical instruments. 

Theorem I asserts that the motion will repeat after a time 2 1 fa. Hence 
if the minimum t period of a vibrating string is determined by observation, 
that minimum period will not be longer than 21/ a. It may be shorter, 
however. For instance, the function 

u(x,t) = sin 2 t rx/l cos 2r at /l 

satisfies (5-1) and the wave equation, hence represents free vibrations of a 
string of length l. But the minimum t period of this function is l/a rather 
than 2 l/a. The shorter period is explained by the fact that uifi/ilfy as 0; 
that is, the center of the string is a node. The center does not move, 
and the string acts like two strings of length 1/2 placed end to end. We 
shall now show that there is always at least one node if the period is smaller 
than that given by Theorem I. 



SEC, 5] THE VIBRATING STRING 445 

Theorem II. If the string considered in Theorem I has an x period 2 p 
or t period 2p/a } where 0 < p < l, then the point x =* p must be a node . 

Suppose, first, that the x period is 2 p. Then, in particular, 


u(p,t) « u(-p,t). 

On the other hand (5-4) gives u(x,l) «• — u(— x,l); hence 

(M) 

«(p,t) - -u(-p,t). 

(6-7) 


By addition of (5-6) and (5-7) we get u(p,t) « 0, which shows that x « p is a node. 
Suppose, next, that the t period is 2 p/a. The equation 



combines with (5-4) to give 

/i(x - at - 2 p) -at - 2p) « /i(x - at) - at). 

If we let x 4 * at * 0 and x — at — 2p ** 8 , the equation reduces, after rearrangement, 
to 

Ms 4- 2 p) - Ms) » c, (5-8) 

where r = /i(0) — f\{ — 2p) is constant. Equation (5-8) shows that /i(s) increases by 
the amount c whenever s increases by 2p. If c 0, it follows that i/i(s) j is unbounded. 
However, fi(s) has period 2/ by (5-5), hence is bounded, and this shows that c « 0 in 
(5-8). The choice s =* —p — at in (5-8) with c =* 0 leads to the desired result: 

n(p,0 ** MV ~ at) - M~P ~ at) ** 0. 

To illustrate the use of Theorem II, suppose a 2-in. -diameter steel cable 
100 ft long is observed to vibrate without nodes at lino rate of two complete 
cycles per second. According to Theorem I the t period is 2 1/a or possibly 
less. But Theorem II shows that the period is not less, since the motion 
was observed to have no fixed points. Hence 

1 100 

- = 2 — 

2 a 

which gives a - 400 fps. This is the velocity with which waves are 
propagated along the cable. Since the density of steel is about 480 lb 
per ft 3 , the weight of l ft of cable is 

480tt(M'2) 2 1 - ( l %)* - 101b per ft. 

This gives p — x %2 slug for the linear density, and hence the tension is 

T « a 2 p - (400) 2 ( 1 5>32) = 50,000 lb. 


PROBLEMS 

1. An infinite string vibrates freely in such a way that the two points x •* 0 and 
x m l remain fixed; that is, u(0,0 « u(l f t ) * 0. Are any other points of the string neces- 
sarily fixed? Which ones? 



446 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

% Suppose a freely vibrating string of length l has just one node at x ** p between 
0 and L Show that the node must be at the mid-point. (The analogous result for n 
nodes is also true.) Hint: If p < 1/2, apply Prob. 1 to the two points x ** 0, x *» p. 
If p > 1/2 , apply Prob. 1 to the two points x ■» p, x «■ l 
8. A cable of length l ft is made of a material with density d lb per ft 8 . It is found 
that the cable makes 10 complete oscillations in t sec. Show that the cross-sectional 
stress is 

a » 0.087ri psi 

provided the oscillations do not have a node between 0 and i IIow w ould the result 
change if the mid-point remains fixed during the observed oscillations but no other 
point remains fixed? 

4. Let h(l) be a given function of t. (a) What is the physical meaning of the boundary 
condition u(l,l ) * h(t)t (b) Describe a physical problem that would lead to the boundary 
conditions u(x,0) * 0, u[x,h(t)\ ** 0. 

6. Initial and Boundary Conditions. We shall now consider the free 
vibrations of a string satisfying the boundary conditions 

v(0 fi = 0, uQ.fi — 0 for all t, (0-3) 

together with the initial conditions 

w(x,0) = /Or), Ui(x, 0) = g(x) for 0 < x < 1. (G-2) 

As in the preceding section we regard the finite string as being an infinite 
string with nodes x ~ 0, x — L According to (5-4) and (5-5), the boundary 
conditions give 

v(: rfi = fi (x — at) — fi(—x — at) (6-3) 

where /x(x) has period 21, and conversely, (G-3) ensures (0-1). The initial 
conditions (0-2) are prescribed on (0,/) for the infinite string, and our task 
is to assign initial conditions outside the interval (0,/) in such a way that 
the solution has the form (0-3). 

Denoting the unknown initial conditions for the infinite string by fo(x) 
and go(x), we have 

/oW = /(*), go(fi = ffW, 0 < x < l, (6-4) 

because the infinite string is to agree with the finite string on (0,1). Upon 
setting t « 0 in (6-3) we get 

fob) */i(*) ~fi(-x). 

Similarly, differentiating (6-3) with respect to t and putting t = 0 give 
Qob) * -~af } (x) + afi(-~$). 

These expressions show that 1 

1 A function tf>(x) wen if <K—x) m 4>(x), odd if <f>(x) m — <f>(x). An analytical and 
graphical discussion of such functions is given in Chap. 2, Sec. 19. 



SEC. 6] 


THE VIBRATING STRING 


447 

( 6 - 5 ) 


/o(x) and go(x) are odd functions. 

Hence, / 0 and go are determined on ( — 1,1) by their values on (0 ,1). 

Finally, since fo(x) and go(x) are expressed in terms of the function fi (x), 
which has period 21, we see that 

/o(x) and g 0 (x) have period 21. (6-6) 

Thus, /o and go are known everywhere as soon as they are known on ( — l, l ). 
According to (3-15), the solution is 

111 rx+at 

u(x,t) * -fo(x - at) + ~/o(x + at) + — / g 0 (s) ds. (6-7) 

2 2 2a 

If /o(x) and g 0 (x) in (6-7) are determined by (6-4) to (6-6), it is easily 
verified that this function u{x,t) satisfies the wave equation, the initial 
conditions (6-2), and the boundary conditions (6-1). Thus, (6-7) is a 
simple and explicit expression for the motion of a vibrating string with 
fixed end points. 

The correspondence between the finite string and the infinite string leads to an 
interesting geometrical construction for getting the disturbance at any point P of the 
strip 0 < t < / in the xt plane (Fig. 10). For the infinite string the disturbance at P is 
found by drawing characteristics as in Sec. 4 (see solid lines in Fig 10). Since the initial 



conditions for the infinite string are obtained from those for the finite string by (6-4) to 
(6-6), the same result may be found by following the dashed lines in Fig. 10. To take 
account of (6-5), however, we must introduce a changed sign upon each reflection at the 
boundary. The disturbance at P arises from the initial disturbance at x{ and * 2 , sub- 
ject to the above-mentioned convention regarding sign. This reflection of the charac- 
teristics in the boundary lines x « 0, x ** / is quite analogous to the reflection of waves 
at the end points of the string. 



44$ 


PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 

The procedure illustrated in Fig. 10 is an example of the method of images, so called 
because the initial conditions for the infinite string are obtained from those for the 
finite string by forming repeated mirror images in the lines t «* 0, x « 0, and x ■* L 
Example: Discuss the free oscillations of a string of length l which satisfies the initial 
conditions 

7l1tX 

w(x,0) * f n sin — - — » tq(x,0) 388 0, 

where « is an integer and /« is constant. 

Since sin nvx/l is odd and has period 21, we may take 


/o(x) « f n sin g 0 (x) ® 0 


as initial conditions for the associated infinite string. Equation (6-7) now yields the 
solution 


u(x,t) 


fn . nir(x — at) f n nr(x -f- at) 

■ 2 sm — 7 + 2 sm — 7 


mrx nired 
® /* sin — cos ■— j~ 


( 6 - 8 ) 


If the initial displacement is given by a Fourier series, so that 1 


u(x,0) «= /(x) ® 2/ B sin — > w,(x,0) - 0 


then superposition of the corresponding solutions (6-8) yields 
u{x,t) 


. Wirx 7} -Kill 

S/ n sm — cos — — 


(6-9) 


Similarly, by choosing /o(x) » 0, yo(x) » pn sin n?rx/f in (6-7), the reader can verify 
that the solution satisfying 

7llTX 

u(x,0) « 0, a<(x,0) « g(r) « 2gr„ sin • 


/ 


/ v Q’J • Hirjr • n7rai 

u(x,t ) * 2* sm — - sm — * 

nira l l 


(6-10) 


Superposition of (6-9) and (6-10) yields the general Fourier-series solution of the wave 
equation satisfying (6-1) and (6-2). The result can be expressed explicitly in terms of 
f{z) and g{x) by means of the Euler-Fourier formulas 

z 2 a \ ' nTX a 2 (' r \ * nirx 7 

fn « y J f{x) sm — dx, g n * y J g(x) sm — dx. 

Because of convergence questions the Fourier-series solution is somewhat less general 
than (6-7), and it is hopelessly inferior to (6-7) for numerical computation. But Fourier 
aeries have great usefulness in that they apply to many problems in which the preceding 

oo 

1 Throughout Chap. 6 we use 2 as an abbreviation for - A brief review of this 

sigma natation is given in Chap. 2, Sec. 2, and Fourier series are discussed in Chap. 2, 
Secs. 18 to 25. 



BBC. 7] THE VIBRATING STRING 449 

methods fail. Examples are given for the vibrating string in the next section and for 
other physical systems in the sections to follow. 


PROBLEMS 


1. Show that the expression (6-8) satisfies the appropriate (a) differential equation, 
(b) initial conditions, (r) boundary conditions. 

2 . The initial displacement of a freely vibrating string of length l is 

„ x 2bx n 1, 

fix) - — > 0 <x <-l, 

fix) m 2b - < X < l t 


and the initial velocity is */(x) * 0. (a) Sketch fix) and foix). (6) Using (6-7) and your 
sketch, find the displacement of the mid-point of the string when at «* 1/4. 

3. (a) Express fix) in Prob. 2 as a Fourier sine series. (6) By (a) and (6-9) show that 
the displacement of the string in Prob. 2 is 


uir,t) 


86/1 ttx Tat 1 . Sir x 3 vat 

(r* s,u T 008 t - p a,n T 008 ~r + • • • 


(c) Obtain an infinite-series representation for the displacement of the mid-point when 
at « 1/4 . 


7. Damped Oscillations. The foregoing discussion was concerned with 
free vibrations, so that F(x r t) == 0 in (2-3). It was indicated that the 
displacement u(x,t) is always periodic in time and hence the amplitude 
remains constant. But. in fact, the oscillations gradually die down when 
a string is vibrating in air, and this behavior is to be analyzed next. 

The reason for the decrease in amplitude is that the air resists the 
motion of an object moving through it. When there is no relative velocity, 
there is no resistance; when there is high velocity, there is high resistance. 
If the resistance is assumed proportional to the velocity, we have 

F(x,t) » —2 bu t (x,t), b > 0, const, (7-1) 

in (2-3). The minus sign is used because the force resists the motion, 
hence is directed opposite to the velocity. Our partial differential equation 
is now 

u it — a 2 u xx » —2 bu t (7-2) 

and the solutions of (7-2) for b > 0 represent the damped oscillations of 
the string. As before, one has the initial and boundary conditions 

u(x, 0) « f(x), u t (xfl) « g(x), (7-3) 

u( 0,0 - 0, u(l,t) - 0. (7-4) 

Equation (7-2) cannot be solved by the method of the preceding sections 
but can be solved by Fourier series. Thus, since the solution u(x,t) is a 



460 PARTIAL DIFFERENTIAL EQUATIONS (CHAP. 6 

twice-differentiable function of x for each t, we may expand u(x,t) in a 
Fourier sine series 

nirx 

u(x,t) =* 2 b n (t) sin * 0 < x < l. (7-5) 

l 


A sine series is chosen rather than a cosine series because such a series 
automatically satisfies the boundary conditions (7-4). To satisfy the 
initial conditions we require 

nirx , nirx 

f{x) « 26 n ( 0) sin — . gix) - 2^(0) sin — • (7-6) 

These relations show that i>«(0) and 6*(0) must be the Fourier coefficients 
of fix) and gix). That is, if 

2 ri htx 2 ft nirx 

/n « 7 / /Or) si n — dr, g n * 7 / flf(x) sin— -dr (7-7) 

l y o l l Jo l 

then multiplying (7-6) by sin nTrrr/Z and integrating from 0 to l yield 

MO) - /«, b' n i 0) = 0n. (7-8) 


We must still satisfy the differential equation. Upon substituting the 
terms (7-5) into (7-2) we get 

k 2 


26" sin — + a 2 26 ; 
l 


■(t) 


T 


— 2625 n sin - 


nirx 

T' 


which gives a set of ordinary differential equations 

„ , / nira \ 2 

b n + 2 bb n + y~J~J b n = 0 


(7-9) 


when the coefficient of sin nirx/l is equated to zero. 

Equation (7-9) may be solved as in Chap. 1 by assuming that b n = e at . 
It is found that 

bnit) = C 0 e~ bi cos u n t + Cie~~ bt sin a? n t, (7-10) 


where 



(7-11) 


The arbitrary constants C 0 and C 1 are determined from (7-8) as 
Co = fny Cl = (gf n + bf n )03~ l . 

Substituting (7-10) into (7-5) yields the final answer 


[fn 


u{x f t) m e 6< 2 fn COS 03 n t + (fif n + bf n ) 


sin u> n l 


w« 


nirx 
sin 

J l 



THE VIBRATING STRING 


451 


SEC. 8] 

Conditions (7-2)~(7-4) are satisfied if the term-by-term differentiation is legitimate, 
for instance, if and g”{x) are bounded. 1 When b « 0, the solution agrees with 
the sum of (6-9) and (6-10), as it should. According to (7-11), the damping reduces the 
frequency of the corresponding terms in the series for undamped vibrations. If 6 < ro/f, 
all the terms are oscillatory and they have the same damping factor e~ bi . But for larger 
values of b the first few harms may have pure imaginary. The corresponding trigo- 
nometric functions become hyperbolic functions, and the terms in question are not 
oscillatory. If w w « 0, which may happen in this latter case, we replace (sin w n Q/wn by 
its limit t [cf. Chap. 1, Sec. 32]. 


PROBLEMS 

1. A string of length l vibrating in air satisfies the initial conditions u(x, 0) 
» fi sin tx/1 , u t (x, 0) ** 0. Show that the displacement of the mid-point can be written 
in the form 

u(l^l,t) ** Ae~~ bt cos (wl -f 4>) t A, eo, <#> const. 

2. Referring to Prob. 1, sketch the curves y « ±Ae~ bt and y « in a single 

neat diagram. Thus describe an experimental procedure for determining b. (When 
the oscillations are rapid and b is small, one can speak of the mean amplitude at a given 
time t. If the amplitude is A o at time to and A \ at time ta r, the reader can verify 
that 



Since Ao and A\ can be found by placing a scale behind the oscillating string, this gives 
a method for comparing the viscosity of gases.) 

8. Forced Oscillations and Resonance. Sometimes the force function 
F(x,l) does not involve the unknown displacement u , as in (7-1), but is 
determined independently. (For example, consider the gravitational force 
on a horizontal vibrating string.) The corresponding mathematical 
problem is 

u tt - a 2 u xx - F(x,t), u(x, 0) = fix), u,(x, 0) = g(x), 

u(0,t) = 0, u(l,t) = 0. (8-1) 

Associated with this problem are two simpler problems, 

v tt ~ a 2 v xx = F(x,t ), v(xfi) = 0, vt(x,0) = 0, 



= 0, 

v(l,t) = 0, 

(8-2) 

and 

w u - a 2 w xz = 0, 

wi(.v,0) = f(x), w t (x, 0) = g(x), 



3 

II 

o 

w(l,t) = 0. 

(8-3) 


Equation (8-2) describes purely forced vibrations, and (8-3) describes free 
vibrations. Now, if v satisfies (8-2) and w satisfies (8-3), it is easily seen 

1 See Chap. 2, Sec. 26, Theorem III. 



PARTIAL DIFFERENTIAL EQUATIONS 


452 


[chap. 6 


that tt *■ v + v> satisfies (8-1). Also, uniqueness in the latter problem 
yields uniqueness in the former. Since (8-3) was solved in Sec. 6, we need 
consider (8-2) only. This system will now be solved formally on the 
assumption that F(x,t) has a Fourier series, 


F(x,t) = 2B n (<) sin (8-4) 

c 

The coefficients are given by the Euier-Fourier formulas, 

2 ri mr£ 

B n(t) -y JT F({,0sin-y^ (8-5) 

Substituting (8-4) and the Fourier series 

MTX 

u(x,t) * 2b n (t) sin ~j~ (8-6) 

into the differential equation (8-1) gives 

„ nirx 0 mrx nwx 

X6 n sin h 2w*b n sin 2 B n sin » 

Z II 

where o>» « nva/l [compare (7-11)]. If we equate the coefficients of 
sin nrz/l , we get 

K + J n b n = B n . (8-7) 

These equations are to be solved subject to the initial conditions 

M 0) - 0, b' n { 0) - 0, (8-8) 

which result from the initial conditions in (8-2). By the method of Chap. 
1, Sec. 28 [cf. also Eq. (33-9) in Chap. 1], the solution of (8-7) and (8-8) is 

Kit) * 0>n l j 0 #n(X) sin Unit - X) dX. (8-9) 

Determining B„(X ) by (8-5), b n (t) by (8-9), and u(x,t) by (8-6) yields an 
explicit formula 

s 2 nirx rt fi nir£ nra 

u{x,t) » 2 sm — / / sm — sin ~~ (t ~ X)F( f,X) d£ dX 

nwa i Jo Jo l i 

when o? n is replaced by its value nira/L 

If we have both damping and forcing, then (8-7) contains an extra term 2bb’ n as in 
(7-9). This leads to a different formula (8-9), but in other respects the analysis is un- 
, changed. Thus, the method of Fourier series enables us to find the damped oscillations 
of a string with arbitrary initial conditions and force function. 



THE VIBRATING STRING 


453 


SRC. 8} 

If F(z,t) is periodic in t, there may be resonance, and this important 
phenomenon will now be discussed for the soecial case (cf. Chap. 1, Sec. 33) 

F(x } t) = a(x) sin cot + b(x) cos uL (8-10) 

[In tlie general case F(z,t) is a sum of terms like (8-10), since the assumed 
periodicity enables us to express F(x,t) as a Fourier series in L] With 
F(xfy as in (8-10) the form of B n (t) can be determined by inspection of 
(8-5). Substitution into (8-7) then gives an equation of form 

b" n + w$6 n = a sin cot + 0 cos u>t, (8-11) 

where a and 0 are constant. 

If a? 2 c* 4 the solutions of (8-11) are all bounded, but if co « w B , the 

particular integral involves the functions 

t sin cot, t cos o)t 

which increase indefinitely with t Hence in that case the term 

riTX 

b n (t) sin — - (8-12) 

in the Fourier series for u(x,t) becomes strongly emphasized as t increases, 
and we say, briefly, that the oscillation (8-12) is resonant. 

A physical explanation is readily given in terms of the results of Sec. 5. 
Thus, the condition o> = co n can be written as 

2tt _ 21 
co na 

This asserts that the period of F(x,t) in (8-10) is equal to the period for 
free oscillations of a string of length l/n. And l/n is precisely the distance 
between nodes for a vibration of the type (8-12). 

Example; A cord stretched between the fixed points x ■* 0 and x *=* l is initially sup- 
ported so that it forms a horizontal straight line. Discuss the oscillations when the 
support is suddenly removed. 

The force function F\(x,t) in Sec. 2 is —gp, and hence the partial differential equation 
is 

Uti — — g , (8-13) 

while the boundary and initial conditions are 

u(0,t) » ** 0, 

u(x f 0) « u<(£,0) m 0. (8-14) 

If wo succeed in finding a particular solution u « v(x) of (8-13) which satisfies the 
boundary conditions 


v(0) m v(l) - 0, 


(8-15) 



454 PARTIAL DIFFERENTIAL EQUATIONS 

then the solution of (8-13) can be written as 


[CHAP. 6 


u(z,t) « w(x t t) + v(x) (8-16) 

where, an follows from (8-13) and (8-14), w(x,t) satisfies 

Wit - oVtx = o, w(0,t) m 0, w(l,t) m> 0, U>(«,0) » — t>(x), 

Wt(xfi) « 0. (8-17) 

Since the desired particular solution v(x) is to be independent of t , the choice u(x,t ) » v(x) 
in (8-13) yields aV' « g, so that 

*>(*) « - ~ z), ( 8 - 18 ) 

when the integration constants are determined so as to satisfy (8-15). This particular 
solution corresponds to the equilibrium position of the string under gravity. The solu- 
tion of the system (8-17) can now be written down with the aid of (6-7) as 

w{x f t) - Hfo(x -at) 4- 34/oOr 4* at), 

where /o(x) is odd, has period 21, and is defined for 0 < x < l by 

fo(x) - -K*) - x(t - x). 

The required solution is u * v 4- w. 

By interpreting /o(-r), fo(x — at), and fo(x 4“ at) graphically one finds that tr(j ,<) is 
largest on 0 < x < l when at * 0, 2/, 41, and then w(x,t) « f 0 (x). Similarly, w{x,t) 
is least when at *» l, 3/, 51, . . and then w(x f t) — —fo(x). It follows that the cord oscillates 
between the horizontal position u » 0 and the position u = 2v{x) in which each point 
is twice as low as the equilibrium position (8-18). The period is 21/ a. 


PROBLEMS 


L A horizontal cable 100 ft long sags 5 ft when at rest under gravity. If the cable 
is disturbed so that it oscillates without nodes, what is the frequency of the oscillations? 
Hint: See (8-18). 

8. A string of length / is subjected to a force F{x,i) ** sin wt sin vx/l, where o> is 
constant. Find the displacement u(x,t) if the string was initially at rest in the equilib- 
rium position. Be sure to distinguish the cases w nra/l and « «* nxa/l. 

3 . Show that the equilibrium shape of a string under a force F(x) is described by 

“-Sjfjf jO?' *>**• 

4 . Show that the function 


v(x,i) 





ds dr 


satisfies v u — a 2 v xx « F(x,t). Hint: Let x 4 at « r, x — at » s, v(x,t) « V(r,s). Then, 
As in Sec, 1, —4 a 2 V r , » F(x,t). 

8. If v(x,t) is the function obtained in Prob. 4, let w{x,l) be determined by 
Wtt — a 2 w xx ** 6, w(x,0) » f(x) — v(x,0), w*(x,0) *» g(x) — v t {xfi) t 



455 


SEC. 9] SOLUTION BY SERIES 

(cf. Sec. 3). Then u — v -f w satisfies 

u tt - a*u zz » F(x,t), u(s,0) » /(x), w t (x,0) - g(x). 

6. Solve the Example in the text by means of Fourier series. 


SOLUTION BY SERIES 

9. Heat Flow in One Dimension. The foregoing discussion of the 
vibrating string enabled us to survey the field of partial differential equa- 
tions and to illustrate a number of important methods. Prominent among 
these is the method of infinite series, which will now be explored more 
fully and used in a variety of applications. We begin with a problem from 
the theory of heat conduction. 

Consider a section cut from an insulated, uniform bar by two parallel 
planes Ax units apart (Fig. 11), and suppose that the temperature of one 


Temperature « w 


u+Au 


x«0 


x x+Ax 


x~l 


Fig. 11 


of the planes is u while that of the second plane is u + Aw. It is known 
from experiment that heat flows from the plane at higher temperature Vo 
that at the lower, the amount of heat flowing per unit area per second 
being approximately 

Aw 

Rate of flow « — k (9-1) 

Ax 

Here A; is a constant called the thermal conductivity of the material; its 
dimensions in the egs system are cal/(em-sec °C). In the limit as Ax 0, 
Eq, (9-1) can be regarded as an exact equality, so that 

Rate of flow = — ku x . (9-2) 

On the other hand, if c is the heat capacity of the medium and p its 
density, the amount of heat in the section from x to x + Ax is 

(cpA A x)u, (9-3) 

where A is the cross-sectional area and w here u is the mean value of u 
over the interval (.r, x + Ax ). For a time interval (f, t + At) the increase 
in amount of heat in the section (z, x + Ax) can be computed from (9-3) 
and also from (9-2). The computation yields 1 

1 It is supposed that no heat is generated within the material and that p f and c 
are constant over the relevant range of temperatures. If p is measured in grams per 
eubic centimeter, the dimensions of c are cal/(g °C). 






466 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

cpA Ax Hix, t + At) — cpA Ax U{x } t) 

* kA At ti-x(x *4* Ax, t) kA At ti x (Xyt)y 

where u x is the mean value of u x in the time interval {t, t + At). Dividing 
by cpA Ax At we obtain 


U( x f t + AO — U(x f t) k u x (x + Ax, 0 ~ u x (Xjt) 


At cp 

and letting Ax — * 0, At — ► 0 now gives 

XX- 1 “• CL Ujj f 


Ax 


.2 « 


Cp 1 


( 9 - 4 ) 


( 9 - 5 ) 


if we recall the definition of partial derivative. 


The fact that (0-4) involves mean values causes no trouble when ut and u xx are con- 
tinuous; see the discussion of Eq. (7-4), Chap. 5. Thus, (9-5) follows without approxima- 
tion from appropriate physical assumptions. This contrasts to the ‘wave equation 
Utt ** which is only an approximate statement of Newton’s law for the vibrating 

string. 


We shall now solve (9-5) under the assumption that the initial tem- 
perature is a prescribed function /(x), 

xi(x, 0) ~ /(x), 0 <x <1, (9-6) 

which can be represented by a convergent Fourier series. The ends of 
the bar are assumed to have the temperature zero: 

u(0,t) * u(l,t) = 0 , t > 0 . (9-7) 

Since u xx must exist if u satisfies (9-5), we know that u(x f t) has a Fourier 
series in x for each fixed t > 0: 

nirx 

u(x,t) = 2 b n (t) sin (9-8) 

l 


Here, a sine series is chosen because such a series automatically satisfies 
the requirement (9-7). Proceeding tentatively, assume that (9-8) can be 
differentiated term by term to give 


, nirx n / riT\ / 

SbUOsin— - « 2 S6 n (0(y) (- 




l ) 


( 9 - 9 ) 


upon substitution into (9-5). Equation (9-9) is satisfied if the coefficients 
of sin nrx/l on each side are equated: 


K 


^arnr 


2 

I K. 



457 


SEC. 0] SOLUTION BY SERIES 

Upon integration this gives 

b n (t) « c n e-<° n ' ll)U , 

where the c„ are constant, and hence (9-8) becomes 

u{x,t) = 2c n e-< an 'l» ' sin • (9-10) 

l 

The initial condition (9-6) yields 

A nirx 

f(x) = u(xfi) * 2c n sin — - • (9-11) 

L 

Since the Fourier series for/(x) converges to f(x) by hypothesis, Eq. (9-11) 
is assured if c n are the Fourier coefficients, 


c n 


2 ri 

T / /(*) sin 
l Jo 


rnrx 

dx. 

I 


The only questionable step in the foregoing discussion was the term-by-term dif- 
ferentiation, but this step can now be justified. Differentiating (9-10) term by term 
actually does give 


u xx 


( nit 

7 


/ 


-lc„e- {a " lty “ (“y y 


(9-1 11 


because the series (9-12) are uniformly convergent when t > 6 > 0. (See Chap. 2, Sec. 7, 
Theorem IV. The uniform convergence follows from the convergence of 

V /{ 2 e -(anril) 2 d^ 

since the Fourier coefficients c„ are bounded.) Hence, (9-10) is a solution of the problem. 
We cannot yet say that (9-10) is the solution, because there might be another solution— 
necessarily different from the one we found - for which the *erm-by-t,erm differentiation 
is not permissible. A uniqueness theorem is established, however, in Sec. 24. 

Because of the exponential factors the series (9-10) is rapidly con- 
vergent and affords a useful means of computing the temperature. By 
contrast, the series obtained in Sec. 0 for solutions of the wave equation 
converges no better than the series for the initial values f(x) and g(x). 
The physical significance of this difference in the two cases is discussed in 
Sec. 27. 

Example 1. Find the steady-state temperature of a uniform bar. 

It is required that u(x,f) be independent of t, whence by (9-5) 

a*u xx ** ut ** 0 . 

Hence, u *■ -h c\x, where co and c\ are constant. If the temperatures at the ends 



458 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

are, respectively, uq and ui, we can determine the constants and thus obtain the formula 

u(x,t) wo -f j (vi — uo). (9-13) 

The rate of heat flow is given by (9-2) and (9-13) as 


and hence, (9-1) holds without approximation in the steady state. 

Example 2. A rod of length 5 has the end x *= 0 at 0°, the end x = 5 at 10 °, and the 
initial temperature is f{x). Find the temperature distribution. 

If v(x,t) is the unknown temperature at point x and time t f we let 

u * v — 2x, (9-15) 

where 2x is the steady-state temperature determined from (9-13). Then a 2 u xx = 
u t) u(x, 0) » f(x) — 2x, u(0,l ) * m( 5,0 *» 0. Hence it is given by (9-10), whore the r n s 
are the Fourier coefficients of f(x) — 2j. When we have found u, Eq, (9-15) gives v 
We have noted that the value 2x introduc<‘d in (9-15) is the steady-state temperature 
as determined by Example 1 The same method enables us to replace anv constant 
boundary conditions by the homogeneous conditions u(0,/) = u(L,t) = 0 That is, if the 
aflknown temperature v(x,t) satisfies 

v(0,t) « % v(l,t) =* vi t v Q and v x const, (9-10) 

we define u to be the difference between v and the steady-state temperature. 

u{x,t) = v(x,t) - [> -b ~ (vi - Vo) J • 

Then u(0,t) *■ u{l,t) * 0, and hence it can be determined by the method of the text. 
A similar use of the steady-state solution was made in the Example, Sec. 8. 


PROBLEMS 

1. Compute the loss of heat per day per square meter of a large concrete wall whose 
thickness is 25 cm if one face is kept at 0°C and the other at 30°C\ Use k = 0.(X)2, 
and assume steady-state conditions. Hint. The wall can be thought to be composed 
of bars 25 cm long perpendicular to the wall faces. By symmetry, no heat flows through 
the sides of these bars in the steady state, and hence (9-14) ran be applied. 

2 . An insulated metal rod 1 m long has its ends kept at 0°C and its initial temperature 
is 50 °C. What is the temperature in the middle of the rod at any subsequent time? 
Use k - 1.02, c - 0.06, and p * 9.6. 

3 . Let the rod of Prob. 2 have one of its ends kept at 0°C and the other at 10 °C. 
If the initial temperature of the rod is 50 °C, find the temperature of the rod at any 
later time. Hint: See Example 2. 

4 . An insulated bar with unit cross-sectional area has its ends kept at temperature 
0, and the initial temperature is f(x) * c n sm nrx/l, where c n is constant and n is an 
integer, (a) Show that the amount of heat present in the bar initially is 2lcpcjnv if 
n is odd and 0 if n is even. ( b ) Show that the net rate of flow out of the bar across the 
ends is 2kc n (nr/l)e^ {anrll)it when n is odd and 0 when n is even. Hint: The rate of 



SOLUTION BY SERIES 


459 


SEC, 10 ] 

flow out of the bar at the end x * 0 is *f ku Xi not —ku x . ( c ) How much heat flows out 
of the bar in the time from t *» 0 to Ct Evaluate a s $ — ► «, compare (a), and explain. 

5. By addition of the results in Prob. 4 obtain similar results for the bar with arbitrary 
initial temperature /(i). 

10. Other Boundary Conditions. Separation of Variables. In the fore- 
going section the differential equation 

Ut ~ CL lixx ( 10 - 1 ) 

was obtained for the temperature u{x,t) of an insulated bar at point x 
and time t. The initial condition was 

u(x, 0) * f(x ), 0 < x < l, (10-2) 

and the ends were held at constant temperature. 

If, instead, the ends are insulated , the boundary conditions are 

uAQ,l) = 0, u x (l,t) « 0. (10-3) 

Equations (10-3) are appropriate because by (9-2) they state that the 
rate of flow across the ends is zero. We shall now consider the problem 
posed by (10-l)-(10-3). 

The boundary conditions (10-3) are satisfied automatically if we express 
u(x,t ) as a cosine series: 

1 71TTX 

u(x y t) = ~a 0 (t) + Sa n (/) cos—* (10-4) 

JL t 

Thus, u x in (10-4) is a sine series (assuming that one can differentiate 
term by term), and we have already noted that the sine series vanishes 
at x ~ 0 and l. 

Substituting (10-4) into the differential equation (10-1) gives 

) a n, (10-5) 

just as in the derivation of (9-9). Solving (10-5) and substituting into 

(10-4), we find 

1 , mrx 

u(x } t) = - c 0 + 2c n e (rm * /7W cos * (10-6) 

2 l 

where the c n s are constant. The initial condition (10-2) shows that the c n s 
are the Fourier cosine coefficients, 

2 ri nrx 

c„ *= - J o f(x) cos — dx, (10-7) 

and the problem is solved. 1 

1 The solution can be verified, if desired, as in the previous section. 


1 ' n ' •if 7 '* 

^ «o 0, (i n ot ^ ^ 



400 


PARTIAL BIFFERENTIAL EQUATIONS 


[CHAP. 0 

We shall now solve this same problem by an important method known 
as separation of variables. It will prove interesting to compare the various 
stages of the solution with the answer, (10-6). 

The desired solution (10-6) is a sum of terms each of which has the form 

X(x)T(t). (10-8) 

In the method of separating variables the idea is to construct functions 
of the form (10-8) which satisfy the differential equation and the boundary 
conditions. By superposition of these functions (10-8), one then satisfies 
the initial conditions. The fact that there is a solution of the type (10-6) 
gives good reason for expecting the method to succeed. 

Substituting (10-8) into (10-1) yields 

XT' = a 2 X"T, 

where the prime denotes differentiation with respect to the appropriate 
variable. Dividing by XT we get 


V 


T 



(10-9) 


The variables x and t in (10-9) are separated , in that the left side is a function 
of t alone and the right side is a function of x alone. It follows that each 
side must be constant, independent of both x and t. A brief investigation 
of the effect of changing sign in (10-10) shows that XT can satisfy (10-3) 
only if the constant is zero or a negative number-* p 2 . Thus, 

rpt 

— * -~p 2 > « ~p 2 . (io-io) 

Independent solutions of (10-10) are 1 

T = e~~ p2i ; X = cos - x, X = sin~x. (10-11) 

a a 


The boundary condition w 2 (0,0 = 0for?4 = XT requires that X'(0) * 0, 
and hence the appropriate choice of X in (10-11) is 


A = cos - x. 


( 10 - 12 ) 


Similarly, the condition u x (l,t) ~ 0 gives X'(t) = 0, so that 

nra 


V 


l 


(10-13) 


*It is suggested that the reader compare XT at this and subsequent stages with the 
general term of (10-6). 



S®0. 10] SOLUTION BY SERIES 461 

where n is an integer. By (10-11)-(1(M3) we see that the function 

T(t)X(x) « e -< nwmll)H cos ~ (10-14) 

satisfies the differential equation and the boundary conditions. To satisfy 
the initial conditions we form a superposition of terms (10-14). The 
resulting series is precisely the series (10-6), and the solution is completed 
as before. 

The merit of the separation method is that it produced the functions 
cos (nrx/l) by direct consideration of the differential equation. If some 
other functions had been more appropriate, the method would have pro- 
duced those other functions instead. This fact will now be illustrated by 
an example. 

According to Newton’s law of cooling, a body radiates heat at a rate 
proportional to the difference between the temperature u of the radiating 
body and the temperature Uo of the surrounding medium. Thus, if our 
insulated rod of length 1 has the end x - 0 maintained at temperature 0 
while the other end radiates into a medium of temperature uq = 0, the 
corresponding boundary conditions are 

n(0,0 « 0, u x (l,t) = — ), (10-15) 

where h is constant. [The second condition (10-15) states that the rate 
of flow — ku x is proportional to u(l,t) — 0, and this agrees with Newton’s 
law.] If h * 0, there is no radiation and we have the condition for an 
insulated end as discussed prc\iouslv. But if h > 0, which we now assume, 
the problem is essentially different from those considered hitherto. The 
difference results from the fact that (10-15) cannot be satisfied in any 
simple way by an ordinary Fourier series. 

Actually, as we show next, the appropriate functions for the problem 
(10-1), (10-2), and (10-15) are not sin (mrx/1) or cos (m rx/I) but arc sin (3 n x, 
where the fi n H are the positive roots of the transcendental equation 1 

(3 cos (31 = ~h sin (31. (10-16) 

Although one could hardly expect to discover the sequence sin fi n x by 
a priori considerations, it is produced automatically by the method of 
separating variables. The solution to the problem is found to be 

u(x,t) = Xc n e~ a2 ^n &ml3 n x, (10-17) 

1 Since the equation is equivalent to tan fil » -0/h when h & 0, its roots can be 
obtained graphically by considering the intersection of the curves y « tan fi and y 
~0/h. Of. Example 2, Sec. 2, Chap 9. 



462 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

where c n is given in terms of the initial values by 

f f(z) sin 0 n x dx 

c n » (10-18) 

/ 8m 2 fi n xdx 

Jq 

To obtain this solution by separating variables, observe that the substitution 
u «» X(x)T(t) leads to functions of the type (10-1 1), exactly as in the former case. Here, 
however, the condition w(0,0 « 0 gives X(0) — 0, so that we require the sine rather 
than the cosine. The resulting expression 

T(t)X(x) - e~ p2 ‘ sin - x 

a 

becomes e~ aSf3 *‘ sin /3x (10-19) 

if we set p * a/3, and this form will be more convenient for our purposes. The function 
(10-19) satisfies (10-1) and the first boundary condition (10-15) for all values of the 
constant 0. To satisfy the second condition (10-15) we must choose (3 so that 

e-aWp co8 _ h € B[n 0 lf ( 10 - 20 ) 

and this leads to (10-16). The resulting functions 

sin (i n x 

satisfy both boundary conditions (10-15) and also satisfy the differential equation 
(10-1). If a suitable superposition (10-17) is found to satisfy the initial condition, our 
problem will be solved 

Setting t * 0 in (10-17) gives 

f{x) « Xc„ sin 0nX. (10-21) 

As in Chap. 2, Sec. 22, Example 1, we can show' that the functions sin & n z are orthogonal 
on (0,/), and hence the r n s are given by (10-18). The solution can be verified by the 
method of Sec. 9 if f(x) admits an expansion (10-21). Since an analogue of Dirichlet’s 
theorem holds for the sequence sin /3 n r, Eq. (10-21) is not a serious restriction on f{x). 


PROBLEMS 

1. If fix) ® g{t), where x and l are independent variables, show that }(x) and g(t) 
are constant. Hint' Let t « to, a fixed value. 

2 . Attempt to satisfy the conditions (10-3) by choosing a positive constant -fp 2 
instead of — p 2 in (10-10). 

3. By using the functions (10-11) solve 

ut * ce 2 u X x, w(0,0 ®= u(l,t) =» 0, w(£,0) ** fix). 

4 . (a) Describe a physical situation which would lead to 

u t « ct 2 u X x, u(0,t) *» u x (l f t) * 0, u(xfl) ** f(x). 

(b) Solve by separating variables [cf. (10-11)]. (c) Verify that your result agrees with 
(10-16M10-18) for h « 0. 

6. Solve Prob. 4 by the method of images. 



SOLUTION BY &EKIES 


463 


SEC. 11] 

Outline of the Solution. Consider a rod of length 21 with ends at temperature 0. Let 
the initial temperature /o(x) agree with f(x) on (0,1), and let / 0 (x) be symmetric about 
x ® l (Fig. 12). By symmetry, no heat flows across the center, and hence the left 




half of the long rod behaves like the rod of Prob. 4. The temperature uo(x,t) for the long 
rod can be found from (9-10). 

6. The vertical displacement u(x,t) of a vibrating string with fixed end points satisfies 

Utt - a 2 o xx , «(0 ,t) = u(l f t) » 0. 

By slitting u(x,t) « X(x)T(t) and separating variables, obtain solutions of the form 

mr at mrx rnrat nxx 

sin — j~ sin — and cos — j— Bin —j~ • 

7. In Prob. 6, express u(x,t) as an infinite series if 

u(xfl) ** f(x), n t (x, 0) « 0. 

8. In Prob. 6, express u(x,t) as an infinite series if 

« 0, u t (x, 0) « g(x). 

11. Heat Flow in a Solid. By a procedure similar to that of Sec. 9 one 
can establish the equation 

k 

U, = a 2 {u XI + u yy + U,,), a 2 -—. (11-1) 

cp 

for the temperature 1 u = u(x,y,z y i) in a uniform solid at time t. This is 
the three-dimensional form of the equation 

u t = at 2 u xx ( 11 - 2 ) 

obtained previously for heat conduction in a rod. The state of the solid 
at time t = 0 gives the initial condition; the state of the surface for / > 0 
gives the boundary condition. For instance, if the surface radiates accord- 
ing to Newton’s law, the boundary condition is 

du 

— = e(u - u 0 ), ( 11 - 3 ) 

dn 

where Uo is the temperature of the surrounding medium, e the emissivity , 

1 See the derivation in Chap. 5, Sec. 16. A similar equation governs diffusion and the 
drying of porous solids, with u equal to the concentration of the diffusing substance. 
Because of this analogy many problems on diffusion and heat conduction are mathe- 
matically indistinguishable. The constant a 2 in (11-1) is often called the diffunvity. 




PAKTIAL DIFFEHENTIAL EQUATIONS 


464 


[CHAP. 6 


and du/&n the derivative in the direction of the outward normal. When 
c*0, Eq. (11-3) means that the body is insulated. 

Sometimes there is so much symmetry that 
u in (11-1) does not depend on y or z . In this 
case (11-1) is the same as (11-2), since the 
terms u vy and u zz in (11-1) are zero, and the 
analysis of Secs. 9-10 can be applied without 
change. 

As a specific illustration consider a uniform plate extend- 
ing from the plane x « 0 to the plane x » d (Fig. 13). 
Let u « uq on the surface x «* 0 and u *» ui on the surface 
x * d, where uo and u j are constant. If the plate is infinite, 
or if the edges are fai away from the points being con- 
sidered, the symmetry suggests that u depends on x only 
and, hence, that (11-2) holds. The steady-state temper- 
ature is then given by Example 1, Sec, 9, as 

x 

u « uo + - (ui - uo). 
a 

Since the rate of flow is —ku x , the amount of heat Q 
flowing across the area A in t sec is 



ktA 


U 0 — U\ 


If the flow of heat is steady, so that u is independent of time, then u t 
and (11-1) reduces to 


u xx + Uy V + u zz = 0. (11-4) 

This is known as Laplace's equation; it occurs in a 
variety of physical problems. The corresponding 
two-dimensional form is 

Uxx + Uyy « 0, u = u(x,y ). (11-5) 

To illustrate the use of (11-5) we shall discuss 
the steady-state temperature in an infinitely long 
metal strip of width d (see Fig. 14). If the sides 
of the strip have the temperature zero and the 
bottom edge has the temperature /(x), the boundary 
conditions are 



Fia. 14 


u(0,y) * 0 , u{d y y) * 0, u(xfi) ** /(x). (11-6) 

We assume besides that (11-5) holds for 0 < x < d, y > 0. 

It is a surprising, fact that these conditions do not suffice to determine 



1 


BBC* 11] SOtUTION BY SERIES 465 

the temperature. 1 However, one expects the temperature to approach zero 
as one moves away from the bottom edge, so that 

lim u(x,y) * 0 uniformly in x . (11-7) 

y — + oo 

If this condition is explicitly required, the solution can be shown to be 
unique (see Sec. 24). 

Although the problem can be solved very simply by Fourier series, we 
prefer to show how the desired functions are generated by the method of 
separating variables. The choice u =* X(x)Y(y) in (11-5) gives 


X" 

T 



( 11 - 8 ) 


after dividing by XF. Since the variables in (11-8) are separated, each 
side is a constant. The boundary conditions applied to XY show (after 
some calculation) that the constant must be a negative number ~p 2 , and 
hence (11-8) gives 



Since (— p) 2 = p 2 , we can assume that p > 0 with no loss of generality. 
Linearly independent solutions of these equations are, respectively, 

cos px, sin px and e vv , c ~ w . 

Since u(0,y) — 0 requires that X (0) ~ 0, we reject the cosine, and in 
view of (11-7) we reject the solution e pv . Hence the function XY takes the 
form 

XY = e~~ vv sin p.r. (11-9) 

The boundary condition u{d y y) ~ 0 gives p = mr/d, where n is an 
integer. Forming a linear combination of the resulting solutions (11-9) 
we get 

, ... mrx 

u(x,y) = 2r n e- (n ’ r/d)1 ' sin (1 1-10) 

d 

and the condition w(x,0) « /(x) now shows that the c n s are the Fourier 
coefficients 

2 rd mrx 

c n “ - / /(*) sin — — dx. 

d J o d 

The solution can be verified, if desired, as in Sec. 9. 

1 The trouble is that the other end of the strip must be taken into account even though 
it is infinitely far away. This purpose is served by (11-7). 



466 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

The foregoing derivation obscures an important point which will now be discussed 
more fully. Although the solutions 

e** and e”™ 

can be chosen for the equation Y " — p 2 Y , these are not the only possibilities. Another 
pair of independent solutions, for example, is 

cosh py and sinh py. 

If, now, we try to decide which of these functions satisfies (11-7) it will be found that 
neither one does. 

What is really involved is the following: The general solution of Y" « p z Y is 

Y * 4. be-rV' 

where a and b are constant. By (1 1-7) we get a =* 0, and hence Y ■ The reader 

can verify that if 

Y ■* do cosh py -f b Q sinh py, 

the condition (11-7) will give ao -f bo *= 0, and again Y is a multiple of e~ pv . Similar 
remarks apply to the construction of X(x) and to the derivation of (10-14). 

Just as in the case of the rod, this problem involving a strip can be given 
a three-dimensional interpretation. That is, the strip need not be thin 
provided there is no variation of temperature across its thickness. By 
letting the thickness approach infinity, wo got a semi-infinite plate. (In 
Pig. 14 the plate extends infinitely far toward and away from the reader; 
the area outlined in the figure is the cross section of the plate, not a frontal 
view.) The boundary-value problem for the plate is 

u xx + u vv + u zz = 0 , 0 < x < d } y > 0, — oo < z < oo, (11-11 ) 

u{Q,y,z) = 0, u(d,y,z) = 0, u(x,0,z ) =/(*), (11-12) 

lim u{x,y,z) = 0 uniformly in x and z. (11-13) 

y — y ao 

If we assume u independent of z , the resulting problem is the same as that 
formerly considered, hence has the unique solution (11-10). 

The fact that u(x,y,z) is independent of z does not follow from the physical symmetry 
but requires the condition (11-13). Indeed the function 

. V® . rry , . 

u «■ sin sin — - e * Tt/a 
d d 

satisfies (11-11) and (11-12) with f(x) = 0 and yet depends on z. Reduction of the 
dimension by omitting a variable is really an application of uniqueness. If we verify 
that (11-10) satisfies the problem (11-11) to (11-13), and that the problem has no other 
solution, then it is true that u must be independent of z. 


PROBLEMS 

1* A refrigerator door is 10 cm thick and has the outside dimensions 60 by 100 cm. 
If the temperature inside the refrigerator is — 10 °C and outside is 20 °C, and if k 0.0002, 



SBC. 12] SOLUTION BY SERIES 467 

find the gain of heat per day across the door by assuming the flow of heat to be of the 
same nature as that across an infinite plate. 

2. If/(x) ** 1 and d » r, show that (11-10) gives 

w«-fc~ v 8inj-f - e~* v sin 3 as *f - sin fir -j V 

it \ 3 5 / 


6. A semi-infinite plate 10 cm in thickness has 
its faces kept at 0°C and its base kept at 100°C. 

What is the steady-state temperature at any point 
of the plate? 

6. The faces of an infinite slab 10 cm thick are 
kept at temperature 0°C. If the initial tempera- 
ture of the slab is 100 °C, what is the state of the 
temperature at any subsequent time? 

7. A large rectangular iron plate (Fig. 15) is 

heated throughout to 100°C and is placed in con- 
tact with and between two like plates each at 0°C. Fig. 15 

The outer faces of these outside plates are main- 
tained at 0°C. Find the temperature of the inner faces of the two plates and the 
temperature at the mid-point of the inner plate 10 sec after the plates have been put 
together. Given: a =* 0.2 cgs unit. Hint The boundary and initial conditions are 

*«M) - 0, ii(3,f) - 0, w(jt,0) - f(x), 

where /(j) * 0 for 0 < x < 1 and 2 < x < 3 but f(x) = 100 for l < x <2. 

12. The Dirichlet Problem. The Laplace equation 

^xx T* T w tz ~ 0 (12-1) 

was obtained in Sec. 1 1 for steady-state heat flow. We shall show how the 
same equation arises in electrostatics and gravitation. 1 

It is a consequence of Coulomb's law that the potential due to a point 
charge q at {?uy\ } Z\) is 

q 

u ~ - taking u = 0 at r = (12-2) 

r 

where r is the distance from the charge to the point (x,y,z) at which u is 
computed. Thus, 

r 2 - (* - Xi) 2 + (y - 2 /i) 2 + (2 - *i) 2 , r > o. (12-3) 

1 A more complete discussion is given in Chap. 5, Sec. 14. The relation of Laplace’s 
equation and fluid flow is developed in Chap. 5, Secs. 15 and 17, and in Chap. 7, Sec. 19. 




468 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

The potential due to a distribution of n point charges g* is given by addition, 

(12-4) 

<- i n 

and the potential due to a distribution of continuous charge of density p 
in a body r can l>e obtained from an expression like (12-4) by passing to 
the limit. 

It is easily shown that 1/r satisfies Laplace’s equation (12-1), and hence 
the same is true of u in (12-4) provided no r t is zero. This latter condition 
means that there is no charge at the point of observation. One would 
expect, therefore, that the potential due to a continuous charge distribu- 
tion will also satisfy (12-1) if there is no charge at the point of observation. 
Tliis is actually the case, and that is the reason why Laplace’s equation 
plays such a prominent role in electrostatics. Although a more sophisti- 
cated treatment may be given, it all comes down to the same thing; namely, 
1/r satisfies (12-1), and the potential is given by some sort of superposition 
process applied to 1/r. 

Since the gravitational potential satisfies (12-2) (where q is the mass of 
the attracting mass point), the study of gravitation also leads to Laplace’s 
equation. In view of its many applications, the Laplace equation (12-1) 
is profitably regarded as a field of study in its own right. Such a study 
leads the way to a branch of analysis known as potential theory . 

An important problem in potential theory is the Dirichlet problem , 
which can be stated as follows: Suppose given a body r in {x,y,z) space, 
together with assigned values /(x,t/,z) on the surface of r. Find a function 
u which satisfies Laplace’s equation in r and is equal to f(x,y,z) on the 
surface. The foregoing discussion gives a number of physical interpreta- 
tions. For instance, if u is temperature, the Dirichlet problem is to find 
the steady-state temperature in a uniform solid when the temperature 
on the surface is given. But if u is the electrostatic potential, the problem 
is to find the potential inside a closed surface when the potential on the 
surface is known. Interpretations in terms of diffusion, fluid flow, and 
gravitation can also be given. 

Since solutions of Laplace’s equation are often called harmonic functions, 
Dirichlet’s problem can be stated as follows: Find a function which is 
harmonic in a given region and assumes preassigned values on the boundary. 
In two dimensions a harmonic function u(x,y) satisfies 

Uxx d~ n yy ** 0. (12-5) 

The region in Dirichlet’s problem is now a plane region, and its boundary 
is a curve. The physical interpretation refers to phenomena in a thin 
plane sheet, or it refers to three-dimensional phenomena which show no 
dependence on z. The latter condition is to be expected when there is 



/ 

SEC. 12J SOLUTION BY SERIES 469 

cylindrical symmetry, that is, when all planes z « const exhibit the same 
geometry and boundary conditions. 

We shall now solve the Dirichlet problem for a circle. It turns out that 
the problem is greatly simplified by use of polar coordinates appropriate 
to the circular symmetry. With 

x * r cos 0, y = r sin 0, u(x,y) m U{rfi) > 

an elementary calculation shows that (12-5) becomes 

(rU r )r + - Uee = 0 (12-6) 

r 

(see Prob. 2). The boundary condition can be expressed as 

U(R,S) - /(0), (12-7) 

where /(0) is a known function of 0 and R is the radius of the circle. 

For each value of r it is clear that U has period 2w in 0, since u is single- 
valued, and therefore U has a Fourier series 

a 0 (r) 

U (r,0) h 2[a n (r) cos nB + b n (r) sin nB]. (12-8) 

2 

Proceeding tentatively, we substitute (12-8) into (12-6) to obtain 

^ + 2,[(ra n ) r cos nd (rb' n Y sin nB] 

2(a n n 2 cos nB + b n n 2 sin nB) ~ 0. 

r 

Since the coefficients of cos nd and of sin nd must vanish, 

( ra' n y = - n 2 a n , n = 0, 1, 2 , . . ., 
r 

(rb' n y « ~n 2 b n , n = 1, 2, 3, .... 
r 

These equations are both of form 

r(ry')' = n 2 y 

which is readily solved by the method of Chap. 1, Sec. 30. Specifically, 
the substitution y « r a gives 

r(ar°y » n 2 r a , 

whence a * dbn. Since a n (r) and 6 n (r) must be finite at r « 0, the minus 
sign is excluded, and 


(? 


<*n(r) * a n r n , 


b»(r) - 6«r“ 



470 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

where a* and b n are constant. Hence by (12-8) 

do 

V ( rfi ) » — + S(a n r n cos n£ + b n r n sin n0). (12-9) 

2 

Putting r « R and using the boundary condition (12-7) give 

do 

f(6) = 1- 2(a n R n cos n$ + b n R n sin nO). (12-10) 

2 


If/(0) has a convergent Fourier series, the validity of (12-10) is ensured by 
choosing a n R n and b n R n to be the Fourier coefficients of /: 

1 rv 

a n R n « - / cos n4> 

ir J ~~ 7r 

(12-11) 

1 n r 

?> n jR n = - / /(</>) sin 7i<f> d<t>. 

7T J — T 


The problem is now solved, but a simpler form can be found as follows: 
Substituting (12-11) into (12-9) gives 


1 n ri /r\" 

w) =-/_[-+ 2 y cos »(•-♦) 


m d<t>, ( 12 - 12 ) 


when we note that 


cos nO cos n<p + sin nO sin n<j> — cos (r/,0 — n<£) 


and interchange the order of summation and integration. The series in 
brackets in (12-12) can be summed as in Chap. 2, Sec. 17, Prob. 0. The 
result is the Poisson formula for a circle 1 


W'-sC 


R 2 — r 2 


2tt •'-» R 2 — 2r/t’ cos (8 — 4>) 4- r 


;/(<*>) d4>. (12-13) 


If /(<£) is piecewise continuous and bounded, one can differentiate under the integral 
sign for r < R to find that (12-6) holds. Also, it can be shown that (12-13) gives 

lira U(rfi) ~ f($) (12-14) 

T — * H — 


provided / is continuous at 0, Hence (12-13) is a solution. In view of the derivation, 
it is remarkable that (12-14) holds even when the Fourier series for/ does not converge 
to/. 

The expression (12-13) gives the steady-state temperature of a thin 
uniform insulated disk in terms of the temperature at the boundary. Or 

1 Another derivation is given in Chap. 7, Sec. 21. 



SOLUTION BT SERIES 


471 


SEC. 13] 

(12-13) can be interpreted as giving the temperature in a circular cylinder 
when the temperature of the surface is /(0) independent of z. On the other 
hand the formula also gives the electrostatic potential in terms of its 
values on the boundary, and so on. 

Example: Let u(x,y) be harmonic in a plane region, and let C be a circle contained 
entirely in the region. Show that the value of u at the center of C is the average of the 
values on the circumference. 

Without loss of generality we can take the center to be at the origin. Equation (12-13) 
then gives, with r — 0, 

“(0,0) - U(0,e) - ~J m d*. (12-15) 

Since /(<*>) stands for the values of u on the boundary, this is the required result. 

PROBLEMS 

1. (a) Verify that 1/r in (12-2) satisfies the Laplace equation (12-1). Hint: rr x *» 
x ~ xi. Using this, find (l/r)*,. ( b ) Verify that logr satisfies the two-dimensional 
Laplace equation (12-5). 

2. If u(x,y) U(r,d) with x « r cos 0, y ■» r sin 0, show that 

Un + Uyv- r~ l {rU r )r + 

Hint: U r * u x cos 0 -f Uy sin 0, U$ * u x (— r sin $) *f u^r cos 0). Similarly, compute 

(rllr)r and (t/,)*. 

3. Derive (12-10) by considering U * R(r) 0(0) and separating variables. 

4. Give two physical interpretations of the following Dirichlet problem for a semi- 
circle, where w(x,y) «*■ U(r,8) as in (12-6): 

Uzx + Uyy - o, X s -f- y 2 < 1, y > 0, 

( 7 ( 1 , 0 ) - g(8), 0 < 0 < », 

r/(r,0) * U(r,t r) - 0, 0 < r < 1. 

5. Solve Prob. 4 by the method of images. Hint: For 0 < 0 < w, define f{9) m g(0) t 
/( — 0) “ —g{8) and use (12-13). 

6. Obtain a formula analogous to (12-13) for the region r > R. (Assume that | U(rfi) j 
is bounded as r — ♦ «o and, hence, that positive values of n in the discussion of the 
text may be rejected.) 

7. Interpret the result of Prob. 6 physically in terms of an infinite metal plate with a 
hole whose edges have a prescribed temperature. 

13. Spherical Symmetry. Legendre Functions. Let it be required to 
determine the steady-state temperature in a uniform solid sphere of radius 
unity when one half of the surface is kept at the constant temperature 0°C 
and the other half at the constant temperature 1°C. By the discussion 
of Sec. 11, the temperature u within the sphere satisfies Laplace's equation / 

u xx + u wv + U B g =» 0. (13-f 

,/ 

1 



472 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

Symmetry suggests the use of spherical coordinates (r,0,4>) with origin 
at the center of the given unit sphere (Fig, 16). Since 

x « r sin 0 cos 

y = r sin 0 sin <£, 

2 ; «» r cos 0, 



Laplace’s equation can be shown to be 1 

r(rU)rr + ( Ue sin 0)o esc 0 + esc 2 0 = 0, (13-2) 

where u(x,y,z) = U(r,0,<t>). If the plane separating the unequally heated 
hemispheres is the xy plane, the symmetry suggests that U will be inde- 
pendent of 4>, so that (13-2) becomes 

T(rU)rr + (Ue sin $)$ esc 0 = 0. (13-3) 

The boundary conditions are 

u = 1 for 0 < 6 < Ylv, when r = 1, 

(13-4) 

u = 0 for < 0 < when r = L 

We shall use the method of separating variables. Substituting the form 

U - «(r)O(0) 

into (13-3) gives two ordinary differential equations, 


r(rR)'' - aR * 0, 
(0' sin 0)' esc 0 + a0 = 0, 



(13-5) 


1 See Chap. 6, Sec. 13, or proceed as in Prob. 2 of the preceding section. 



BOLtmON BT SERIES 


473 


BBC. 13] 

where a is an arbitrary constant. The first of these equations can be Bolved 
by assuming that R — r"* as in Chap. 1, Sec. 30. One obtains the linearly 
independent solutions 

R « r m , R = r“ (m+1 \ 
where m satisfies the quadratic equation 

m(m + 1) = a. (13-6) 

Changing the independent variable in (13-5) from 0 to x by mejtnw of 
X = cos 0, 0(0) = P(x), 

and replacing a by the expression (13-6), we get Legendre’s equation 

(1 - x 2 )P" - 2 xP' + m{m + 1)P = 0, ' = — • (13-7) 

dx 

When m is a nonnegative integer, a solution of (13-7) is the Legendre 
polynomial P m (x) = P m (eos0). Thus, one is led to consider solutions 
of (13-3) which have the form 

r m P m (cos9) or r^^P^cos 6). 

The second of these expressions is rejected because it becomes infinite 
as r — > 0, and we attempt to build up the desired solution u by forming a 
series 

GO 

« = 23 A m r m P m (cos 0). (13-8) 

0 

Each term of this series satisfies (13-3). 

When r = 1, Eq. (13-8) becomes 
00 

« = 23 A m P m (CO8 0), r *= 1, (13-9) 

m«*0 

and if it is possible to choose the constants A m in such a way that (13-9) 
satisfies the boundary condition (13-4), then (13-8) will be a solution of 
the problem. Since x = cos 0, the boundary condition requires 

CO 

F(x) = 23 A m P n (x), (13-10) 

m*»0 

where F(x) ® 0 for —1 < x < 0, and Fix) * 1 for 0 < x < 1. Now, it 
was stated in Chap. 2, Sec. 22, that the expansion (13-10) is possible for 
suitably restricted functions F{x) and that the coefficients are gi 

= ^ + j_ x F(x)P m (x) dx. 




[chap. 6 


474 PARTIAL DIFFERENTIAL EQUATIONS 

By means of this formula, the solution is found to be 

U~H + HrPli cos e) - K6r 3 P 3 (cos 6) + %^ 6 (cos #)-•••■ 

It is possible to establish that (13-8) is actually a solution, though the demonstration 
requires a detailed knowledge of Legendre functions. 1 The uniqueness theorem es- 
tablished in Sec. 24 shows that there is no other solution, and hence the foregoing 
procedure can be justified. In particular it was permissible to take m as a nonnegative 
integer and to use the polynomial solution of (13-7) rather than one of the infinite-series 
solutions. 


PROBLEMS FOR REVIEW 

1. As an infinite series, express the steady-state temperature in a circular plate of 
radius a which has one half of its circumference at 0°C and the other half at 100°C. 

2 . By (12-13), find the temperature of the plate considered in Prob. 1. 

8. By separating variables in polar coordinates find the steady-state temperature in 
a semicircular plate of radius a if the bounding diameter is kept at the temperature 
0°C and the circumference is kept at the temperature 100 °C. 

4 . Interpret the following Dirichlet problem physically, and solve: 

Uxx *f ■“ 0, 0<3C<l,0<y<l, 

u{0,y) rn w(l,y) - u(x,0) m 0, «(*,1) - f{l). 

5. Derive (13-5) from (13-3). 

14. The Rectangular Membrane. Double Fourier Series. Let a uni- 
form elastic membrane be stretched 
over a fixed, plane, bounding curve 
(Fig. 17). To explain what is meant 
by the tension , we consider the force 
AF exerted by the membrane on one 
side of a small straight slit of length 
As. The membrane is said to be 
under uniform tension T if this force 
is directed perpendicular to the slit in 
the plane of the membrane and has 
magnitude T As independent of the 
location and orientation of the slit. 
A similar definition applies when 
the membrane does not lie in a plane except that we must let As — * 0: 



Ai o As 


The role taken by the plane of the membrane in the first case is now taken 
t>y the tangent plane at the point in question. 

1 One must show that the series obtained by differentiating (13-8) are uniformly con- 
vergent f or r < 1 — 5 and that the boundary condition is verified as r — ► l. 




SOLUTION ST SBBIBS 


SBC, 14] 


475 


Let the coordinate system be so chosen that the bounding curve of the 
membrane lies in the xy plane. The vertical displacement of any point in 
the membrane at time t is denoted by u — u(x,y,t). To obtain a dif- 
ferential equation for the motion, we consider a small, nearly square portion 
of the membrane bounded by vertical planes through the points 


(s,y,0), ( x + Ax, y , 0), (x, y + Ay, 0), (x + A x, y + Ay, 0) 

(see Fig. 18). Applying Newton's law to the small portion gives the ap- 
proximate equation 

T 

u H =* y 2 (u xx + Uyy), y 2 * — t (14-1) 

P 


where p is the surface density. This equation describes small oscillations 
of the freely vibrating membrane. Its derivation is similar to the cor- 
responding derivation for a vibrating string (Sec. 2). 



The problem of the vibrating membrane is solved when we have found 
the solution of (14-1) which satisfies appropriate initial and boundary 
conditions. We shall now consider the case of a clamped rectangular 
membrane with sides of lengths a and b (Fig. 19). The boundary conditions 

art ' u ® 0 for x *= 0 and for x = a, 0 < y < b, 

(14-2) 

u * 0 for y *= 0 and for y ** b, 0 < x < o. 



To determine the solution uniquely we also specify the initial displacement 
and initial velocity: 

u{x,y,Q) - f(x,y), u t (x,y, 0) = g(x,y). (14-3) 



478 


PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 


The assumption that 
in (14-1) yields 


t* - X(x)Y(y)T(t) 



(14-4) 


upon division by XYT. Since the variables are separated, the terms in 
(14-4) are constant. It can be shown that these constants are negative, 
so that we may write 

X" 2 Y" T" 2 

se — » —a 2 , sm — (*), 

X Y T 

with y 2 (p 2 + y 2 ) * o> 2 by (14-4). 

Since X H + p 2 X ** 0, the function X(x) is a linear combination of 
sin px and cos px. The cosine is rejected because the condition u ~ 0 at 
x =» 0 gives X(0) = 0, and we must have p = tmr/a , where m is an integer, 
because the condition u « 0 at x * a gives X(a) = 0. In just the same 
way it is found that 

Y = sin yy, 

where # * rvw/b for an integer n. Thus, the desired oscillation has the form 


mrx niry 

sin sin {A cos o> mn t + B sin u mn t), 

a b 

where A and B are constant and where w m « = a? is given by 



(14-5) 


The functions (14-5) satisfy the differential equation and the boundary 
condition. To satisfy the initial conditions (14-3) we try a superposition, 
using different constants A and B for each choice of m and n: 


~ mwx nrry 

u(x y y f l) “ JL C 4mn COS o) mn t + B mn sin Wnnt) sin sin — — (14-6) 

m,n— >1 U 5 


Since the initial displacement is f(x,y), we must determine A mn so that 

/(*,») 


mxx niry 
2, A*,* sin sin—— 

wi.n—l O 0 


Multiplying this double Fourier series by sin (vhtx/q) sin {nry/b) and inte- 
grating over the rectangle give the formula 


4 


•b 


mrx 



SOLUTION BY SERIES 


477 


SEC. 14 ] 

just as in the corresponding discussion for single Fourier series (Chap. 2, 
Sec. 18). Similarly, differentiating (-14-6) with respect to t and setting 
t « 0 give 

4 r* rb mrx niry 

Bmn - -7 / / Bin sm — dx dy 

abu> mn J ° Jo a b 


when we use the second initial condition (14-3). 

The general term of the series (14-6) is a periodic function of time with 
period 2ir/<*w The corresponding frequencies 


Wmn 

~2t 



cps 


(14*7) 


are called characteristic frequencies , and the associated oscillations (14-5) 
are called modes. The fundamental mode is the mode of lowest frequency, 
obtained by setting m = n = 1. 

Similar terminology applies to the vibrating string (Secs. 2-6). If the 
length of the string is a and the equation of motion is 

u tt « y 2 u xx , 


the characteristic frequencies may be written in the form 



(14-8) 


analogous to (14-7). The modes are described by 


mrx 

u = sin (^4 cos co m t -f B sin u) m t) 

a 


and the fundamental is the mode obtained for m « 1. The three-di- 
mensional analogue of (14-7) and (14-8) is discussed in Prob. 2. 

In Sec. 8 it was shown for the vibrating string that the characteristic fre- 
quencies agree with the resonant frequencies, and a similar behavior is 
found for vibration phenomena in general. It is also true in general that 
the vibration can be expressed as a superposition of individual modes. 
This fact is illustrated by (14-6) and by the Fourier-series solution for the 
vibrating string. 

The behavior of the vibrating membrane differs from that of the string in one respect. 
For each characteristic frequency of vibration of the string the corresponding mode is 
such that the string is divided into equal parts by the nodes whose positions are fixed. 
When a membrane oscillates with a given characteristic frequency, there are also points 
on the membrane which remain at rest. Such points form nodal Hrm . The position 
and the shape of the nodal lines, however, need not be the same for a given frequency. 



478 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

As an illustration consider a rectangular membrane with a — b, The frequency equa- 
tion (14-7) then yields 

tamn -“Vm*Tn* 

a 

« aVm* Hh n* « » — • (14-9) 

a 

For m » n «■ 1, we get from (14-6) the fundamental mode 

Uu — (An cos 4* Bn am o n t) Bin — sin — > 

a a 

where «n «■ o\/2. Since ttu « 0 for all t only when x ® 0, y « 0, x « a, y « a, there 
are no nodal lines in the interior of the membrane for this frequency. If we take m » 1, 
n ** 2 and m *» 2, « =» 1, we get two modes: 

rX , 2ir^/ 

*» (-4 is cos 4- Bn sm wiaf) sin — sm » 

a a 


2ttZ try 

t*si - (4 si cos a>sif + Bji sin a> 2 i<) sin — sin — » 

a a 

with the same frequency, since u> 2 i «= u>u = a\/5. For y * a/2, t * A 2 * 0 and for x * 
a/2, n«i ** 0. These nodal lines are shown in Fig. 20. By forming linear combinations of 
the modes in (14-10) we can get oscillations with the same frequency but with different 
nodal lines. Thus, if we take An * An « 0 and form uu 4- uu, we get 




</I5 


at*** /I? 


n»=cot /IS 









BBC. 14] 


SOLUTION BT SERIES 


479 


tin 4* u®i m sin uud ( Bn sin — sin ----- + Bjj sin — sin 

\ a a a a/ 

— (sin o»nf)2 sin — sin — ( Bn cos h B 2 i cos — 1 • 

a a \ a a / 

For this oscillation the nodal lines in the interior of the membrane are determined by 

« ~ irx 

Bn cos h Bfl cos — — 0. (14-11) 

a a 

Equation (14-11) for Bn * #21 yields the nodal line z 4- y * a and for Bn * -Bj 1 , 
the line x — y «** 0 (see Fig. 20). Different nodal lines can be obtained by forming dif- 
ferent linear combinations of the modes (14-10). 

The reader will show that for m * n «* 2, all oscillations have the same nodal lines 
x » a/2, y « a/2, while infinitely many different nodal lines can be obtained by form- 
ing different linear combinations of the modes iq* and un. A few of these are shown in 
Fig. 20. 

Since the nodal lines may be regarded as the boundaries of new membranes contained 
in the original one, the character of oscillation of membranes of different shapes can be 
deduced from the examination of nodal lines (see Prob. 3). 

Nodal lines can be observed experimentally by sprinkling a fine powder on the vibrat- 
ing membrane. 

PROBLEMS 


1. Suppose the initial conditions for the rectangular membrane considered in the 
text are 

u(x,y, 0) — 0.1 sin ~ sin u t (x,y,0) « 0. 

a b 


(a) What is the frequency of the oscillation? (b) What is the maximum opeed attained 
by the mid-point of the membrane? 

2 . Analysis of a microwave resonant cavity leads to the equation 

u t t — y +yw -f u«) 

with the boundary condition u ** 0 or du/Jn « Oon suitable portions of the planes 
£* 0 , x**a, j/ 0, y «* b, z *» 0 , z — c 


(see Fig. 21). By assuming u ** XYZT show that the 
characteristic frequencies are 


Wmnp 

IT 


[©'♦©♦©T 


where m, n, and p are integers. 

3. A curve in the xy plane along which u — 0 for all t is 
called a nodal line . (a) Sketch the nodal lines for the oscill- 
ation (14-5). (6) Sketch the nodal line for the oscillation 


( . rx . 2ry . 2*x . ry\ . 

am — sm ~ — h sm sm — ) (A cos ws\t 4- B sm a*i0 

a 6 a b / 



Fia. 21 


which arises by adding the modes m — 1, n ** 2 and m ■* 2, n * 1. Hint ; sin 26 » 
2 sin & cm 0. (c) Thus obtain one solution for the problem of a triangular membrane. 



PARTIAL DIFFERENTIAL EQUATIONS 


{chap, 6 

16 . the Circular Membrane, Bessel Functions. To discuss the oscil- 
lations of a circular membrane with fixed edges we introduce cylindrical 
coordinates ( r,$,u ). With 

u(z,y,t) « U(rfi t t) (15-1) 

the equation of motion (14-1) takes the form (cf. Sec. 12, Prob. 2) 

V U - y 2 [U r r + T~~ l U r + T~~ 2 Um], (15-2) 

If the boundary is the circle r ~ a, then the boundary condition is 


t7(a,0,<) = 0. (15-3) 

To make the problem definite we also introduce initial conditions 

U(r,0fl) = /(r), Ut(r,$fl) - 0 (15-4) 


which state, respectively, that the initial shape of the membrane is given 
by f(r) and that the initial velocity is zero. 

Since the initial shape is independent of 6 , the solution presumably in- 
volves r and t only. Thus, we consider expressions of the form R(r)T(t) 
When applying the separation method. Substituting into (15-2) gives 


1 d 2 T ~ 2 /l d 2 R 1 dR\ 
TlF ~ y \Rdl + 7Rd^) 


(15-5) 


after division by RT . Since the left-hand member of (15-5) depends on t 
alone and the right-hand member on r alone, each side must be constant. 
It can be shown that the constant is not positive, hence may be written 
as — o 2 . Thus (15-5) leads to 


+ 

So 

II 

© 

/ f 

dt 

(15-0) 

R" + r~ l R' + k 2 R = 0 , 

d 

' * — . 
dr 

(15-7) 


where k « v/y. 

Equation (lS-fi'l is the familiar equation for simple harmonic motion, 
and Eq. (15-7) can be reduced to the Bessel equation by the substitution 
x » kr. Hence, (15-7) has a solution 

R 581 J q{x) *** </o(&r). 

The other solutions of (15-7) are rejected because they become infinite 
at r * 0, and we are led to the functions 

Jo(kr) sin wl or Jo(kr) cos uL 

Since u t * 0 when t m 0, we reject the solution involving the sine. The 



SBC. IS] 


SOLtTTION BY SERIES 


481 


boundary condition (16-3) applied to our elementary solution RT now 

Jo(ka) cos u>t 0 


for all t. This requires that ka be a root of the equation Jq(x) = 0 (see 
Fig. 22). If the positive roots of Jq(x) are denoted by x n , the appropriate 



choices of k are given by 
tions have the form 


k n ~ x n /a. Sir.ce w = ky, our elementary solu- 
RT = Jo(k n r) cos k n yt. 


These functions satisfy the differential equation (15-2), the boundary 
condition (15-3), and the second initial condition (15-4). To satisfy the 
first initial condition we try to represent U as a linear combination of such 
terms: 

U » 2 A n J 0 (k n r) cos k n yt. (15-8) 

n=*l 

When l * 0, the initial condition requires that 

00 

f(r) = E A n J 0 (k n r). 

«=* 1 

The problem of expanding an arbitrary function in series of Bessel functions 
was discussed in Chap. 2, Sec. 22. It was shown that the coefficients are 
given by 2 r« 

<1M) 

provided the series is uniformly convergent (but see also Chap. 2, Sec. 23). 

In the terminology of the preceding section, the solution (15-8) is ex- 
pressed by means of the modes. The characteristic frequencies are 

Wn Ky x n y 
2r 2r 2ra 

and the fundamental is described by Jo(k x r) cos k x yt. 



PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 


482 

PROBLEMS 

1. The oscillations of a cylindrical resonant cavity satisfy 

uu « y\ux* + %+ u Zjf ), 0 < r < a, 0 < * < 6 

with boundary condition u * 0 on the curved surface, u z ** 0 on the plane ends. Obtain 
solutions of the form R(r)Z(z)T(t ) for this problem, 

2. Find the distribution of temperature in a long cylinder whose surface is kept at 
the constant temperature zero and whose initial temperature in the interior is unity. 

3. An elastic membrane subject to uniform gas pressure satisfies the equation 

u it + V * 7 2 (w« + %*), 

where p is a constant depending on the pressure. If the membrane is circular, show 
how to reduce this problem to a problem of the type solved in the text. Hint' Consider 
the function 

C7(r,9,0 


SOLUTION BY INTEGRALS 

16. The Fourier Transform For many partial differential equations 
the desired solution can be expressed as an integral involving the initial 
or boundary values. This possibility was already illustrated by formula 
(3-10) for displacement of a vibrating string and by the solution of the 
Diriehlet problem given in (12-13). We shall now describe a systematic 
method of obtaining integral formulas. 

The function g(s) defined by 

1 fa 

T f * lim ■ := / e lzs f(x) dx * g(s) (1G-1) 

q — 4 op “y/ 2w * 

is called the Fourier transform of f(x) ; the operator T is called the Fourier 
transform operator. The inverse operator T _1 is obtained by changing 
the sign of i, so that the foregoing equation may also be written 

1 f a 

I J= lim — 7 == / e' x ’g(s) ds = f(x). (16-2) 

When such is the case, the symbol T satisfies the easily remembered equa- 
tions 

TT* 1 / - f, T-'T/ - /. (16-3) 

If the limits in (16-1) and (16-2) are regarded in the sense of mean convergence (Chap. 
2, Sec. 23), and if the integrals are regarded as Lebesgue integrals (Appendix C), then 
(16-1) gives (16-2) and (16-2) gives (16-1) provided either of the integrals 



BBC. 16] SOLUTION BY INTEGRALS 483 

is finite. 1 Both integrals (16-4) then have the same value. In many physical problems 
the common value represents the total power or energy present in the system. 

To illustrate the use of the Fourier transform, we shall solve the problem 


U| - a 2 u xx , 

t > 0, —00 < X < 00 , 

(16-5) 

s — s 

i! 

o 

£ 

— 00 < X < 00, 

(16-6) 

u(z,t) — > 0 , 

as t — > oo. 

(16-7) 


Physically, this system describes the temperature u(x,l) of an infinitely 
long bar at point x and time t when the initial temperature u(xfi) is known. 
The trial solution u » e px + 9t with p and q constant leads to 

qe px + qt « a 2 p 2 e p *+ qt 

when substituted into (16~5). Hence q * a 2 p 2 > and the trial solution is 

e 1>x+a*p l t' 

We choose p 2 negative because of (16-7). Thus p « is, where s is real, 
and the trial solution is now 

e i,x-a'.H = c i,x e -a',*t. ( I (>-8) 

We shall satisfy the initial condition (16-6) by forming a linear combina- 
tion 2 of solutions (16-8). Thus 

_L e *»* e - a, 'Vs) 

\/2ir 

is a solution of (16-5) no matter what value g(s) may have, and the integral 
u(z,t) = -J== f e % * x e-* iait g(8) ds 

V 2x •'—oo 

is also a solution, provided we can differentiate under the integral sign. 
By (16-2) the latter expression can be written 

u(x,t) = T -‘e— “**Vs). (16-9) 

Setting t = 0 and using the initial condition (16-6) give 

fix) = T - 1 0 (s), 

1 This important theorem, known as PlanckereVs theorem , is proved in E. C. Titch- 
marsh, “Introduction to the Theory of Fourier Integrals,” Chap. 3, Oxford University 
Press, London, 1937. For a heuristic discussion of the relation between (16-1) and 
(16-2) see Chap. 2, Secs. 20 and 21, of the present text, 

* This procedure is analogous to the formation of Fourier series in the method of 
separating variables. 



484 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 8 

so that Tf « TT ~~ l g ** g. Substituting into (16-9) we get the final answer, 

u(x,t) « (16-10) 

This is an explicit formula for the temperature u(x,t) in terms of the 
initial temperature /(x). 

As another example we shall solve the Dirichlet problem for a half 
plane. Several physical interpretations were given in Sec. 12; the mathe- 
matical formulation is 


"1" Uyy 

= o, 

y > 0 , —oo < x < oo, 

(16-11) 

u(x, 0) 

= fix), 

— 00 < X < 0O y 

(16-12) 

u(x,y) 

- o, 

as y — * oo. 

(16-13) 


The function e px+9V satisfies (16-11) if p 2 + g 2 = 0. We choose q real 
and negative because of (16-13), and hence p is pure imaginary, p = is. 
The trial solution is now 

e ) 

when we note that g 2 = $ 2 and that q is negative. This function satisfies 
(16-11) and (16-13). To satisfy (16-12) we form a linear combination as 
in the previous example, thus: 

1 

u(x,y) = -~= / ds m T ~ l e~ u ^g. 

v 2 tt * —so 

Setting y = 0we get / = T -1 g by (16-12). Hence g = T/, and the solu- 
tion 1 is 

u(x,y) * (16-14) 

As a final example we shall consider the problem 

u tt = a 2 u xx , u(x, 0) = /(x), u t (x, 0) = g(x) 

which describes waves on an infinite string (cf. Secs. 2 and 3). The trial 
solution e ux + qt yields the two expressions 

e ™ x e iaat an( J e i*x e ~i<i8t t 

Forming a general linear combination, we get 

u(x,t) ~ T~V a ‘Vi (s) + (16-15) 

where g\ and g 2 are to be determined from the initial conditions. 

If (16-15) can be differentiated under the integrals which are implied 
in the result is 

* T^uue^ffiis) - ( 16 - 16 ) 

1 Formulas of the type (16-10) and (16-14) are discussed in R. M. Redheffer, Operators 
and Initial-value Problems, Proc. Am, Math. Soc. t 4 (August, 1953)* 



485 


SEC. 16] SOLUTION BY INTEGRALS 

Setting t ** 0 in (16-15) and (16-16) now gives, respectively, 
f = T ~ l gi + T“Va» 0 “ T -1 *os0i — T~ 1 iasg a 
or, after operating on the equations with T, 


01 + 02 

Solving for Q\ and g 2 , 

1 


T/, 108(0! - 0 2 ) = T0. 


01 


■Tf + —~Tg, 

2 2 xcls 


1 1 

02 - - V — — - T0, 

2 2tas 


(16-17) 


and this gives the final answer upon substitution into (16-15). 

The foregoing result can be deduced from d’Alembert/s formula (3-15). 
However, the method of Fourier transforms also applies when d’Alembert’s 
method fails (cf. Probs. 1 and 2). 


PROBLEMS 


1. According to Sec. 7 the equation for damped motion of waves on a string is 

u t t — a z u** <■ —2 bu t . 

Obtain a family of solutions of the type 

«(*,<) - + T- 1 e-* , e-‘( 6, -» v >^j(s) 

by starting with u ** e l,x + qi and forming a linear combination. 

2. Formulate appropriate initial conditions for Prob. 1, and use them to determine 
gi and pa- 

3. The displacement u(x } t) of a long, stiff rod satisfies 

EI ^ mc * m f ° rCe ’ 

when the mass is negligible (cf. Sec. 2, Prob. 6). Let U(s,t ) be the transform of u with 
respect to the variable x, and let F(s,t) be tho transform of /. Neglecting convergence 
questions, show that EI* A U *» F, and thus obtain the solution in the form 

u(x,t) - (A’A)- 1 T- 1 s*~ 4 T/. 


Hint: Write out the expressions 

u(x f t) - T-'UM, 


f(x,t) - T - 1 F(bA 


in full, and substitute into the differential equation. 

4. If the mass of the rod in the preceding example is m, the equation of motion is 


d 4 u d 2 u 
El 4" wi - 


dx 4 


dt* 


■toO- 


Show that u “ T~ l f7, where U ■» J7(s,0 satisfies the ordinary differential equation 





486 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 8 

17, Waves to a Half Plane* The Fourier transform can be used to solve 
the two-dimensional wave equation 

y 2 (u xx + Uyy) * u Ui y = const, (17-1) 

and the result has an interesting physical interpretation. We suppose 
that the time dependence is harmonic, so that 

u(x,y,t) - V{xa,)e-**, (17-2) 

where <*> is constant. Substituting in (17-1) gives the scalar wave equation 

U xx + Uyy + k 2 U ~ 0, k - -• (17-3) 

7 

This equation will now be solved in the half plane y > 0 subject to the 
additional conditions 

U(z,0) - f(x) } U(r 9 y) -> 0 as y «. (17-4) 

Physically, the solution describes the radiation field of an antenna 1 when 
the aperture illumination is /(r) (see Fig. 23). 



By substituting the function e %xsJtqv into (17-3) we obtain solutions 

c .« e ±i (17-5) 

Because of the second condition (17-4) the coefficient of y in (17-5) has a 
negative real part when s is large; we shall indicate this by dropping the 

1 Formulation of (17-3) in this context and discussion of the conditions for a unique 
solution can be found in treatises on electromagnetic theory. 



SBC, 17] SOLUTION BY INTEGRALS 487 

minus sign. Forming a linear combination of expressions (17-5) as in the 
preceding Section, 

U{x,y) - — j= r da « 

y/2* •'-» 

For y * 0 the first condition (17-4) yields g = T/, and hence 

U(x,y) - T-^^T/. (17-6) 

Multiplying by e*-" 1 "* as in (17-2), we get the corresponding solution of 
(17-1). 

The discussion of (17-1) given here contrasts to that given in connection with the 
vibrating membrane (Sec. 14). For the membrane we specified the initial values of 
u and u t , and we obtained a series involving infinitely many oscillation frequencies. 
Here, on the contrary, the frequency «/2» was prescribed in advance. By (17-2) the 
initial conditions are 

- U(z,y) t utfayfl) « -iuU(x,y) 

and the first condition cannot be specified arbitrarily, inasmuch as U (x,y) satisfies (17-3). 
To interpret the solution physically, we have 

u(x,y,t) « T~ l e~ iut+ ' v '^~* l g(s) 

by combining (17-2) and (17-6), with T f *= g(a). Writing out in full, 
u(x,y,t) - ds . ( 17 - 7 ) 

For simplicity we shall suppose that g(s) = 0 when \s\> k. The limits 
(-- 00 , 00 ) of the integral can then be replaced by (— k t k). If we now in- 
troduce a variable B , 

8 « k sin 6, y/k 2 — $ 2 *» kcosd, (17-8) 

we get 

u(x,y,t) - f rl2 e* (ix *+* •— °fl(fc sin 0)fc cos 0 d9. 

\/2w 1 *73 

This formula expresses the solution as a superposition of functions of the 
form 

sin 6+kv cos 0— (17-9) 

In the next paragraph it will be seen that (17-9) represents a plane wave 
traveling with velocity y in the direction 0 (Fig. 23). Hence, the Fourier 
transform procedure gives the plane-wave expansion of the antenna field . 
The amplitude of the wave moving in a given direction 0 is 

g(k sin 0) k cos B dB } 

where p(s) is the Fourier transform of the aperture illumination. 



488 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 8 

To see that (17-9) represents a plane wave we examine the (x,y) locus on which (17-9) 
is constant* Without essential loss of generality we confine our attention, in particular, 
to the locus along which the exponent is zero. The equation of this locus is 

x sin $ 4* y cos 6 ** ~ t «■ yt, (17-10) 

when we divide by k and replace k by its value (17-3). 

The wave front (17-10) is clearly a straight line, and the wave fronts for different 
ts are all parallel. In fact, their common perpendicular makes an angle $ with the y 
axis. Since the distance from the line (17-10) to the origin is yt, the velocity of propaga- 
tion is the constant y, and hence the desired result is established. It is possible to give 
a similar discussion for the part of the integral (17-7) with |s| > k, though we shall not 
do so here. 

18. The Convolution Theorem. The convolution f * g of two functions 
/ and g is defined by 

f*g = lim ~ f fii)g(x -{)<#£ = - 4 = / /(f)fif(x - {) (18-1) 

if the limit exists either in the ordinary sense or in the sense of mean con- 
vergence. The importance of the operation (18-1) rests on the following 
theorem: 

Convolution Theorem. Let f, g, |/| 2 and \g\ 2 be integrable, and let 
all infinite integrals be interpreted in the sense of mean convergence. Then 
the product of the transforms equals the transform of the convolution. In 
symbols, 

T (f*g) « (T/)(T?). (18-2) 

Although a complete proof requires knowledge of Lebesgue integration and mean 
convergence, the result cao be made plausible as follows We have 

TO. ,) . — f_ - 1) *] d. 

provided the order of integration may be inverted. If the variable x in the inner integral 
is changed to t « x ~ £ , we get 

k £ [ fj^ mu ■“] **> « - (vs £ ,<8 *- «) (vs £*■>•-“ 4 

and this is (Tf)(Tg). 

By means of the convolution theorem some of the foregoing results 
can be greatly simplified. Taking the transform of the formula (16-10) 
with respect to x gives 

Tu = e-“ v< T/ 


(18-3) 



I 

BBC. 18] SOLUTION BY INTEGRALS 488 

for the temperature u(x,t) of a rod when the initial temperature is f(x). 
By consulting a table of Fourier transforms or by using the result of Prob. 
1, 

e = T g(x), where g(x) = (2a 2 <)“ H e-*' / < 4 **«. 

Hence, the result (18-3) may be written 


Tu =» (Tf)(Tg) = T(f* g) 

when we recall (18-2). Taking the inverse transform now yields 

u(x,t ) = f * g — (4 xa 2 t)~^ f /(f)e” ( * _f), ^ 4a,<) df. (18-4) 

' —00 


The advantage of this formula is that it involves only a single integration 
whereas (16-10) requires two integrations. Since the integral is rapidly 
convergent when t is not too large, (18-4) is well suited for numerical 
computation. 

To obtain a physical interpretation of (18-4), let the rod have the initial 
temperature zero except for a short piece on the interval (x 0 — xq + «) 
(see Fig. 24). If Q cal of heat is uni- 
formly distributed over this element |.2*| 

of the rod, the corresponding initial s — 1 

temperature / is given by *o 

Q ** 2 ecpf, Xo~~€<X<Xo + € ^ IG ’ ^ 

where c is the heat capacity and p the linear density. By (18-4) the resulting 
temperature at point x and time t is 


u(x,t) 


- I r° + ‘ <,-(*-{>*/ (4a*t) d t 

cp(4xa 2 2) ^ 2e •' x o~« 


Letting « — » 0 and using the mean-value theorem we get 

9 g— (* — * 0 ) */ (4o*t) ^ 

Cp(4:ira 2 t) ^ 


(18-6) 


This gives the temperature distribution for an instantaneous source of 
strength Q at the point x 0 . Now, Eq. (18-4) represents the temperature in 
the general case as a superposition of such sources . The source at z = { 
has the strength 

Q 33 cpf(Z) d£. 


Am we shall see in the next section, this physical interpretation enables us 
to solve a variety of problems in heat flow with the greatest ease. 



400 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

Example: Let u(x,y) be harmonic for y > 0 and satisfy the additional conditions 
t<(x,0) — f{x), u(x,y) -* 0 as y —► *>. 

Show that u w given by the Poisson formula for a half plane: 

u(x,y) - ? f f -~rr~ t di. (1M) 

x (x - £)* + y * 

Since this problem is the same as that in (16-11) to (16-13), the solution is given by 
(16-14). Taking the transform of (16-14), 

Tu « e-^nj. (18-7) 

The convolution theorem can be applied if we express c~ ,,lv as a Fourier transform. 
To this end we compute the inverse transform 


T~ l c~W* - ~4 f= r *” x e-‘ v d8 / e^e'vds 

y/2r Jo \/2 r J ^ 

+ JLV 

\/2x \y — ix y -f xx) 


a&$) 


This shows that «■ Tjr, where g is the function (18-8): 

1 2 y 

9 “Vz*** + v* 

The convolution theorem applied to (18-7) now gives u «= /* g, and that is the desired 
result. 


PROBLEMS 

1. Let l(x) - T _1 c“ e#2 , where c is constant, (a) Differentiate, and integrate the 
result by parts to obtain 

dJ _ x 
dx 2c 

(b) Using Eq. (10-1), Chap. 8, find the value of 1(0). ( c ) Thus deduce the formula 

T -i e ~e.* m (2 c)~ H e-* lic . 

(In particular, e~ x * f2 and e~ fl/2 are transforms of each other.) 

2. Obtain the temperature distribution for a rod extending from x • 0 to x ** <*> 
if the initial temperature is /(x) and the end x * 0 is insulated. Hint: Consider a rod 
extending from — *> to w, with initial temperature fo(x) defined by 

fo(x) even, /o(z) « /(x) for x > 0. 

Compare Prob. 5, Sec. 10. 

& By taking fo(x) odd in Prob. 2, find the temperature distribution when the end 
x — 0 is not insulated but is kept at the temperature zero. 

4. A rod extending from x *» 0 to x «* l has the initial temperature distribution /(x). 
By regarding this rod as part of an infinite rod with the initial temperature /o(x), find the 
temperature u(x,t) when (a) both ends are insulated, (6) both ends are kept at the tem- 
perature zero. Hint: Let /o(x) have period 2 L This method of satisfying boundary 
conditions was used for the vibrating-etring problem in Sec. 6. 



t 


SEC. 191 SOLUTION BT INTEGRALS 491 

19. The Source Functions for Heat Flow. According to (18*5) the 
function 


V e -r*/ 
pc(4*0L 2 t) ** 


x 2 , 


(19*1) 


represents the temperature distribution due to an instantaneous source 
of strength Q at the origin. Equation (19-1) applies to the one-dimensional 
heat equation c?u xx = u t . The corresponding result for two dimensions 
is 

__0 -r W) r* - X s + y 2 , (19-2) 

pc(4iror0 

and for three dimensions it is 


V e ~r*/4 

pc(4ira 2 0 ^ 


r 2 » x 2 + y 2 + r 2 . 


(19-3) 


In these formulas r is I/ie distance from the source to the point of observation 
and t is the length of time that has elapsed since the heat was released. The 
value of p is, respectively, the linear, surface, or volume density. 


The functions (19-1) to (19-3) are solutions, respectively, of 

a 2 U*x « U t> a 2 (u*r + Uyy) « Ut, a\u xx + Uyy ' f U Mt ) - U t . 

Also they give the limit 0 as l —► 0 through positive values, provided r ^ 0. Hence 
the initial temperature distribution is concentrated entirely at the origin. By integrating 
over the whole space it can be shown in each case that the total amount of heat present 
is Q when t > 0. For these reasons, the physical interpretation as a 'point source oj 
strength Q is fully justified. 

The expressions (19-1) to (19-3) indicate that heat travels with infinite speed. Even 
if r is large, we get a positive temperature for each positive t, no matter how small, but 
the initial temperature was zero. By contrast, the disturbance associated with the 
wave equation travels with finite speed (cf. Secs, 2 and 4.) 

To illustrate the use of (19-3) let us find u(x,y,z,t) when the initial 
temperature 

u(x x ,y x ,z x> 0) = f{x h y u zi) 


is given at each point (x x ,y x ,z x ) of space. Instead of this distribution we 
introduce a source of strength 

Q = cpf(x x ,y x ,z x ) dxi dy x dz x (19-4) 

at (x x ,y x ,z x ). The temperature at point (x,y,z) and time f due to one such 
source is given by (19-3), with Q as in (19-4) and with r the distance from 
(x,y,z) to the source: 

r 2 - (xi - x) 2 + ( y x - y) 2 + (z, - z) 2 . 



PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

Hie temperature at (x,y,z) due to all the sources is given by superposition: 
u(x,y,z,t) « (4*a 3 t)~ H [ f f e~ rVia, 'f(xi,yi^i) dx r dy t dz x . ( 19 - 5 ) 

J — oo J — co J — oo 

As another illustration we shall find the temperature due to a point 
source which emits heat continually. Let Q(t) represent the strength of 
the source, so that the amount of heat emitted in time interval (ty, ty + dty) 
is approximately Q(ty) dty, The heat at the present time t due to the source 
at time U is nn \ /it 

pc[4rra 2 (t — ^i)]^ 

when we recall that t in (19-3) stands, not for the time, but for the elapsed 
time. Adding the contributions from the source at all values of ty prior 
to t gives 


u 


-f 

DC 




pc [4?ra 2 (£ — ty)]** 


- ,l) 1 Qih ) dh 


(19-6) 


If Q(t) is a constant Q, the integral can be evaluated explicitly by the change of 
variable 

r 2 


the result is 


u « 


4o ! « - h) 1 

Q I 

4x0? pc r 


G&"7) 


This represents the temperature due to a continuous uniform source of heat at a distance 
r from the point of observation. Since the conditions are steady state, the solution 
satisfies Laplace's equation. (Compare Sec. 12, where the function 1/r was obtained 
in connection with electrostatics and gravitation.) 

Example: A line contact is pressed against the 
plane x ** 0 with constant normal force F per 
unit length, the coefficient of friction being a 
constant p. At time t ** 0 it starts to slide in 
a direction perpendicular to its length with con- 
stant velocity v (see Fig. 25). Obtain the tem- 
perature in the medium x < 0, assuming this 
temperature to have been zero initially and 
neglecting heat loss at the surface x « 0. 

This problem arises in the theory of milling, 
leather glazing, and lathe turning. To solve it, 
let the line contact be initially coincident with 
the z axis, so that its height at time ^ is y «■ vty. 
The heat generated by friction per unit length 
is Fp dy, and hence the heat generated per unit 

Q mm Fpvdty, 

Using this value of Q in the result (19-2) we obtain 



e -l**+ /4a»(i-l,) 

- h) 



SOLUTION BT INTEGRALS 


493 


sec. 20] 

for the contribution, at the point (x,y,z) and at the present time t, due to motion of the 
line contact in the time interval (h, h -f dti). (The reader is reminded that t in (19-2) 
stands for the elapsed time and x 2 -f y 2 in (19-2) is the square of the distance from the 
point of observation to the line source.) Superposition yields the final answer: 

- — y o — dh. 


PROBLEMS 


1. Show that (19-2) can be obtained by integrating (19-3) with respect to z and (19-1) 
by integrating (19-2) with resect to y. Interpret physically. 

2. What initial- or boundary-value problem is solved by (19-5)? By (19-6)? 

3. By use of (19-2) solve the initial- value problem 

oVx* 4 * %) « u u u(x,y, 0) « /(x,#). 

4. Find the temperature distribution u(x f y,z y t) for x > 0 due to a time-dependent 
distribution f(y t *,t) on the plane x ■» 0. Take the initial temperature as zero for x > 0 

6, State and solve the two-dimensional analogue of Prob, 4; the one-dimensionaJ 
analogue. 


20. A Singular Integral. We shall now derive an integral formula which 
can be used in the study of many partial differential equations. The 
discussion depends on certain theorems of vector analysis 1 summarized 
in the following paragraph. 

In the divergence theorem 

/(V.A)dv«/A n da, 


dr =* volume element 
da = surface element 


(20-1) 


the choice A *» uVv yields Green's first identity 


r 0 r dv 

/ [uV 2 v + ( Vn • Ur;)] dr = / u — da 
Jr J<r dn 


(20-2) 


when we recall that (Vt>) n — dv/dn, the normal derivative. Writing 
(20-2) with u and v interchanged and subtracting give Green's symmetric 
identity 

r n n r ( dv du\ 

I ( uV 2 v — vV 2 u) dr » / ( u v — ) da. (20-3) 

J r J o \ dn dn/ 


The conditions for validity of these identities are discussed in Chap. 5. 
For our present purposes we need an appropriate form of (20-3) when 
v does not satisfy the continuity conditions there required. 

1 The reader may find it advisable to review Chap. 5, Secs. 8 to 10. Unless otherwise 
indicated, the functions considered in Secs. 20 to 22 are twice continuously differentiable 
in the region r and on its boundary. The surfaoe of r is assumed smooth, so that the 
normal is a continuous function of position. 



494 


PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 


To this end, we show that 

f M) , f4n KP), fore = 2, 

a -* o An a e 10, for c < 2, 


(20-4) 


where / is continuous, c is constant, and the region of integration or is 
the surface of a sphere of radius a centered at P. Now, the integral (20-4) 
may be written 


An a® An a e A» 


/(P) 


/(C) -/(P) 


da = /i + 


Since the area of the sphere is 4ira 2 , we have, as a — ► 0, 


/i 



a c 


4?ra 2 — > 


1 4ir/(F), 

to, 


for c = 2, 
for c < 2. 


Since a surface integral does not exceed the area of the surface times the 
maximum value of the integrand, we have 


r , _ 2 max |/(Q) -/(P) | 

/ 2 1 < 4 wtr 


<4* max |/(Q) -/(P)|. 


If / is continuous, this tends to zero as a — ► 0, and 
(20-4) follows. 

Let us now apply (20-3) to the function 
1 

v m w + r « r(P,Q), (20-5) 

r 

where w is twice continuously differentiable and 
where r is the distance from a fixed point P to the 
variable point of integration, Q . The region of 
integration is to be the region inside a given closed 
surface a and outside a small sphere a\ of radius a 
centered at P (see Fig. 26). In this region r & 0 
and (20-3) can be applied without hesitation. 

According to (20-5) we have 

V 2 v - + V 2 - » V 2 w. (20-6) 



On a i the outward normal n is directed along the radius into the sphere, 
so that 

- 1 * _ _ I 1 

dnr dr r r* a 2 



SEC. 21] SOLUTION BY INTEGRALS 495 

Since r ® a on <ri, the foregoing equation and (20-5) give 
Bv Bw 1 1 

— 1 — v « w + -» on (T\. (20-7) 

Bn Bn or a 

The surface integral in (20-3) can be written as an integral over a plus 
an integral over crj. By inspection of (20-7) the integral over vi is 

f r / Bw i\ / i\dui 

+ (2M) 

'ITiis becomes 4iru(P) as a — ► 0, in view of (20-4). If we use this result 
and (20-6) in (20-3), we obtain the desired formula 

4 ru(P) = f ( uV 2 w — vV 2 u ) dr + f (v u — \ da (20-9) 

\ Bn Bn/ 

upon letting a —> 0. When P is exterior to r, the same formula is valid, 
except that 4 wu(P) must be replaced by 0. 

Since the volume integral in (20-9) is taken over the whole region r, it includes the 
point P at which v «* *>. The meaning of the integral is clear from the derivation, 
but we shall show directly that a singularity of the type 1 fr in a volume integral causes 
no convergence difficulties. If n is the interior of the sphere with surface c\ t we have 

f - dr « f - 4 rr 2 dr » 2rd 2 . 

Jti r Jo r 

This is clearly finite and in fact tends to zero as a — ► 0. 

21. The Poisson Equation. If u has continuous second derivatives, then 
the Laplacian V 2 u is a continuous function of position. We shall denote 
this function by —4 irp(z,2/,z), so that 

V 2 w = —4 t p. (21-1) 

The choice u * 1/r, te « 0 in (20-9) now yields the Poisson formula 

r P 1 r /I Bu B 1\ 

u(P) ~ -dr + -/(--- u-Ada (21-2) 

r 4 v \r Bn Bn rf 

when we divide by 4ir. As before, r * r(P,Q) is the distance from P to 
the variable point of integration Q . The formula (21-2) holds for every 
function having continuous second derivatives in r and on its boundary. 1 

We now change our viewpoint. Instead of starting with u and defining 
p by (21-1), we suppose that p is given in advance. Equation (21-1) 
is now a partial differential equation for the unknown function u; it is 

1 Provided the boundary is simple enough to permit the use of (20-9). This condition 
on r is hereby postulated once for all. 



496 


PARTIAL DIFFERENTIAL EQUATIONS 


(CHAP. 6 

called the Poisson equation . The foregoing considerations show that if u 
satisfies the Poisson equation, then u is given by the Poisson formula. 
The interest of the formula is that it yields the values of u throughout the 
interior of r in terms of u and du/dn on the surface only. 

For a physical interpretation, let u be the electrostatic potential due to a 
charge distribution of density p. The fact that the potential satisfies 
Poisson's equation is established in treatises on electrostatics, 1 so that this 
interpretation is consistent with (21-1). Since g/r represents the potential 
due to a charge q at a distance r from the point of observation, the term 

1 

- (p dr), where r = r(P,Q) and p = p(Q), 
r 

represents the potential at P due to the charges within the volume element 
dr at Q. Hence the first term of (21-2), 



represents the potential at P due to charges within the body r. Similarly 
the second term in (21-2), 



represents the potential at P due to a certain surface-charge distribution 
on the surface a. 


To interpret the term 

/(£;)( -£■<■) < 2M » 

in (21-2), we consider the configuration shown in Fig. 27. Here, a charge — g is in- 
troduced at the point Q on the surface <r and a charge +q at a distance An along the 
outward normal n to <r. The distance from —q to P is r, and the distance from q to P 
is denoted by n. If we take q « m/An, where m is constant, the potential at P is 

1 1 1 A(l/r) d 1 

q q — ** qA - » m ► m 

ri r r An dnr 

as An — ► 0. [That the limit is m d(l/r)/dn follows from the definition of normal deriva- 
tive, without calculation.] The limiting configuration of Fig. 27 is called a dipole; 
the constant m is called the moment of the dipole. We have thus found the desired 
interpretation of (21-3); namely, (23-3) represents the potential due to a surface distribu- 

1 The case p - 0 in (21-1) is discussed in Chap. 5, Sec. 14. A detailed analysis of the 
conditions under which Poisson’s equation holds may be found in 0. D. Kellogg, “Foun- 
dations of Potential Theory," p. 156, Springer-Verlag OHG, Berlin, 1929. 



SEC* 21] SOLUTION BY INTEGRALS 497 

tion of dipoles having the moments ~u de/4r. A surface distribution of dipoles such 
as this is called a double layer . 

Since the volume integral in (21-2) is extended only over r, it does 
not take account of the charges outside r. That purpose is served by the 
surface integral in (21-2). From this viewpoint (21-2) shows that the 
chargee outside r can be replaced by a suitable surface charge and double 
layer on <r, without changing the potential within r. If r increases beyond 
all bounds, the limiting value of the surface integral can be thought to 
represent the influence of the charges at infinity. 




In many important problems there are no charges at infinity, so that the 
limiting value of the surface integral is zero. To investigate this possibility, 
let <r be a large sphere of radius a centered at the origin 0. By inspection 
of the differential triangle in Fig. 28, 


Ar ^ An cos ~ An cos as An — ► 0 

where ^ is the angle between OQ and PQ. The definition of normal deriva- 

tive leads to 

dr 4 Ar 

— — lim — = cos ip 
dn An 

(21-4) 

and hence 

d 1 1 dr CQ8\p 

dn r r 2 dn r 2 

(21-5) 


If P is fixed and a it is easily seen that r ~ a uniformly with 

respect to the point Q on a. Hence 1/r has the order of magnitude 1/a, 
and by (21-5) the normal derivative has the order of magnitude 1/a 2 . 
Now, the surface integral in (21-2) does not exceed 4xa 2 times the maximum 



498 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

of the integrand. By the foregoing remarks, the integral therefore tends 
to aero as a *> if 

— ► 0 and max \ u\ -* 0, as a — > », (21-6) 

In this case (21-2) leads to the simple formula 

u « f - dr, integrated over all space. (21-7) 

* r 

Referred to spherical coordinates 1 (r,6 f 4>) the normal derivative du/dn 
in (21-6) is the radial derivative du/dr , and a == r. Thus (21-6) is equiva- 
lent to 

du 

lim r — » 0, lim u * 0, uniformly in 6 and (21-8) 
r rn dr r — ► «o 

By substituting (21-5) into the integral and regrouping terms one finds that (21-8) 
may be replaced by the weaker condition 

( du \ u 

r — -f- u ) « 0, lim - » 0, uniformly in $ and <t>. (21-9) 

dr / r — ► so r 

That is, if u satisfies (21-1) and (21-9), then u can be represented in the form (21-7). 
When p » 0 outside a bounded region, an altogether different procedure 1 shows that 
the second condition (21-8) also suffices. 

Example: If the region r is a sphere of radius ro centered at JP, every solution of Pois- 
son's equation satisfies 

U<P) " l “ da + 1 ( ; _ “) ' dr - 

Here r » r(P,Q) is the distance from P to the variable point of integration Q. To 
prove (21-10) we choose w — —1/ro in (20-5) and note that on a 

dv dv 1 l 

P "" # dn dr r* ™ rj> 

The desired result follows at once from (20-9). The special case p » 0 in (21-10) yields 
the Average-value Theorem: If a function is harmonic throughout a sphere , its value 
at the center of the sphere equals the average of the values on the surface . This fact is of 
central importance in the study of harmonic functions. 

The merit of taking w - — l/r 0 in (20-5) is that then v *» 0 on «r. Hence the term 
involving du/dn in the surface integral (20-9) drops out. The possibility of making 
such a choice of v will be systematically exploited in Sec, 23. 

1 The r in (21-8) has no relation to the r m r(P y Q) that appears elsewhere in this 
discussion. 

•See H. B. Phillips, "Vector Analysis/' p. 158, John Wiley <fc Sons, Inc., New York, 
1083. 


a max 


du 

dn 



sbc. 22] 


SOLUTION BT INTEGRALS 


t 


PROBLEMS 


1. If u ia harmonic, show that the choice u ■» c in (20-2) give* 

jf + «?) dr - jf tt ^ dr. 

2 . Show that & solution of Poisson’s equation in a closed bounded surface <r is wholly 
determined by its boundary values and that it is determined, apart from an additive 
constant, by the boundary values of the normal derivative. Hint: If tq and u% are two 
solutions, apply Prob. 1 to u «* u\ — u* and then use the result of Prob. 4. 

8. Let u be harmonic in a region r, and suppose u assumes its maximum value uq 
at an interior point P. Show that u ** uo throughout every sphere contained in r and 
centered at P. Hint : If M(u) denotes the mean value on the surface of such a sphere, 
then uo ■* M(u) and hence M(uo — u) «■ 0. Now use Prob. 4. 

4, Let f(Q) be continuous and nonnegative in a region r. If J f dr » 0, then/ m 0. 

Similarly for surface integrals. Hint: If / • « > 0 at an interior point P, then by 
continuity / > e/2 throughout some sphere r\ of radius 6 > 0 centered at P. But this 

gives J f dr > J f dr > jf («/2) dr > 0. 


22. The Helmholtz Formula. The Helmholtz equation 

V*u + k 2 u - 0 (22-1) 

is obtained by separating variables in the wave equation (Sec. 14) cr by 
requiring harmonic time dependence (Sec. 17). A brief calculation 1 shows 
that (22-1) has the solution e %kt /r i where r is the distance from a fixed 
point P to a variable point Q. If we set 

e xkr 1 

v * — » w =*= v (22-2) 

r r 

it follows that V 2 w » V 2 v =* —k?v for r 0. Hence 

uV 2 w — vV 2 u » u(—k 2 v) — v(—k 2 u ) = 0, (22-3) 


provided r 0 and provided u satisfies (22-1). Substituting (22-2) into 
(20-9) with due regard to (22-3) now yields the Helmholtz formula 


m 


i 

4w 


*• L dn dn V r / J 


da. 


(22-4) 


Tlsl* expresses tbfi solution u of (22-1) as an integral involving the boundary 
va%WB of u and du/dn. 

Sometimes the region r is bounded and (22-1) holds at points exterior to 
t. To see if (22-4) remains valid in this case, we construct a sphere t\ 


1 Let the Laplacian be referred to spherical coordinates with origin at P, 



PARTIAL DIFFERENTIAL EQUATIONS 


[CHAP. 6 

centered at the origin and having a radius a so large that r is contained 
entirely within n (Fig. 29). Formula (22-4) applied to the region between 
t and the surface <r\ of the sphere gives 


rfe^du B /e ikr \l r rv* 

4ru(P) «/ u — ( — )\d<r + I — 

r dn dn \ r / J r 


^ r du 
dn 



d<r , 



In the first integral the outer normal for the region of integration is the 
inner normal for r. With this understanding we see that (22-4) holds in 
the present case, provided the integral over cr x tends to zero as a — * 

To investigate the behavior as a -* », note that on <r\ 

d e* r d dr e** r / 1\ 

r* w ( lk I COS ^ 

dn r dr r dn r \ rj 


by (214). Hence the integral over <r\ becomes, after rearrangement, 



— iku + iku(l — cos f ) + - u cos J cUr. 


(22-6) 


As a — ► oo with P fixed, we have a ~ r. Also, the law of cosines applied to Fig. 28 
shows that o 2 (l — cos ^) remains bounded. Hence, the integral will tend to 0 provided 


a max 



0 and max | w | — ► 0 f as <*—►». 


This assumes k real, so that |« ftr | ** 1. 

In spherical coordinates (r,6,4>) the result of the foregoing discussion 
may be summarized as follows: Formula (22-4) applies to the exterior oj 
the hounded region r provided 


r 


du 

,dr 



0 


(22-6) 


and w 0 



SOLUTION BY INTEGRAL 


SEC. 28] 


SOI 


as r — > », uniformly in 6 and 4>. In just the same way it is found that 
(21-2) applies to the exterior of the bounded region r provided (22-6) holds 
with A: ■» 0. 


Equation (22-6) with km 0 is sometimes called the Dirichlet condition . It is the 
same as (21-8), hence means that there are no charges at infinity. Equation (22-6) 
with k & 0 is called the Sommer f eld radiation condition; it means that there are no 
sources of radiation at infinity. 

Although (22-6) is the form usually given, it is unnecessarily restrictive. A more 
careful analysis of (22-5) shows that (22-6) may be replaced by the weaker condition 


( du u\ u 

— — iku *f ~ 0 and - — » 0, (22-7) 

which reduces to (21-9) when k — 0. 

23. The Functions of Green and Neumann. The Laplace equation can 
be obtained by setting p *= 0 in Poisson's equation (21-1) or by setting 
k = 0 in the Helmholtz equation (22-1). The corresponding integral 
formulas, (21-2) and (22-4), both reduce to 


<P) 



d 1\ 

u ] do. 

dn tI 


(23-1) 


This expresses every harmonic function in r as an integral involving the 
boundary values and the boundary values of the normal derivative. How- 
ever, a harmonic function is determined by the boundary values alone , 
without any reference to the normal derivative. 1 We shall now obtain a 
formula, similar to (23-1), in which du/dn is not present. 

Such a formula can be found by an appropriate choice of v in (20-9). 
Since V 2 u » 0, the volume integral in (20-9) will drop out if 

V 2 w * 0 throughout r. (23-2) 

And the terra involving du/dn in (20-9) will drop out if v «= 0 on <r. By 
(20-5), that condition is equivalent to 

w » on a. (23-3) 

r 

Evidently, (23-2) and (23-3) determine w uniquely. Since the function 
r as# r(P,Q) involves the fixed point P, the boundary condition (23-3) 
makes w> and hence t\ depend on P. The function v obtained in this way 
is called Green 1 8 function and is denoted by G(P,Q). Thus, 

1 

G(P,Q) « v ** w + - 

r 


(23-4) 



PARTIAL DIFFERENTIAL EQUATIONS 


502 


[chap. 8 


where r ** r(P,Q) and where w satisfies (23-2) and (23-3). 
(20-9) now yields 


u(P) 


1 r dv If 

— I u — d<T w / u 

4% dn 


dG 


4r ** dn 


d«, 


The formula 
(23-5) 


with G given by (23-4). The differentiation and integration in (23-5) are 
with respect to Q. 

What we have shown is the following: Let u satisfy 

V 2 w - 0 in r, u » / on or. (23-6) 

If the region r has the Green function G, then 

1 r dG 

U (P) * f f(Q) — da . (23-7) 

4* dn 

When a continuous function / is given in advance, it can be shown, con- 
versely, that the function u in (23-7) satisfies (23-6). In other words, 
formida (23-7) solves the Dirichlet problem. The general Dirichlet problem 
is thus reduced to the special Dirichlet problems 1 that have to be solved 
in constructing Green’s function. 

To interpret Green’s function physically, let a unit charge be placed 
at the point P interior to a closed, grounded conducting surface a. Since 

P is the only charge present, the potential 
has the form v =» w + 1/r, where V 2 w — 0. 
Since the conductor a is a grounded equi- 
potential, v = 0 on cr, and hence, v agrees 
with the v in the foregoing paragraph. 
Thus y G(P } Q) is the potential at Q due to o 
unit charge at P in the grounded conducting 
surface <r. Because of this interpretation 
the existence of Green’s function is verj 
plausible on physical grounds.* 

The physical interpretation not only suggest 
that G(P,Q) exists but gives a method of findini 
it in many cases. As an illustration, we shal 
construct Green’s function for the half space z > t 
(Fig. 30). Let a charge q « 1 be placed at P and a charge q ** —1 at Pi, the mirrc 
image of P in the plane * * 0. By symmetry the potential v ■» 0 when t « 0, an 
hence v is Green’s function. If r is the distance from P to Q and n the distant 
from Pi to 0, the potential is 

1 One problem for each choice of P. 

* A proof of the existence for all regions likely to be met in practice is given in O. I 
Kellogg, "Foundations of Potential Theory,” chap. II, Springer-Verlag QHG, Berlii 
1929. 




fisc. 23] 


SOLUTION BY INTEGRALS 


SOS 


0(P t Q) ----- 
r n 

As in the derivation of (21-5), 


dO 

dn 


2 cos \ff 


on * *» 0, 


where ^ is the angle between PQ and the normal to * *» 0. Substituting into (23-7) 
yields the Poisson formula for a half space , 


u(x,y t z) 



cos ^ 

_ 


d<r 


* r r i(x\,y\) . 

2t J-^ [(* - x,) 1 + (y- vtf + i*J* ^ 2,1 


This formula represents a harmonic function for z > 0, which reduces to f(x,y) when 

t *■ 0 . 


In terms of heat flow, the Dirichlet problem is to compute the steady- 
state temperature in a solid when the temperature on the surface is known* 
Sometimes the rate of heat flow across the surface is prescribed rather than 
the temperature. The problem which arises in this way is called the 
Neumann problem; it leads to the equations 


V 2 u « 0 in r, 


du 

— = <7 on a. 
dn 


(23-8) 


If (23-8) is to have a solution, we must restrict g so that the rate of 
flow into r equals the rate out; otherwise, a steady-state temperature 
cannot be expected. It is clear physically that the appropriate condition 
is 

/ g da - 0 (23-9) 

J<r 

and indeed, the choice v » 1 in (20-3) shows that (23-9) follows from 
(23-8). 

When g satisfies (23-9), the problem is still not well posed because it has 
infinitely many solutions. That is, (23-8) involves the derivatives only, 
so that u can be altered by an additive constant. To make the solution 
unique we require that 

£ u da « 0. (23-10) 

Properly stated, the Neumann problem, is to solve (23-8) when (23-9) and 
(23-10) hold. 

By means of (20-9) we can develop a Neumann function N(P,Q) ana- 
logous to the Green function G(P,Q ) of the foregoing paragraphs. As 
before the condition 


V 2 w « 0 throughout r 


(23-11) 



804 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 0 

makes the volume integral (20-9) drop out. To get rid of the surface 
integral involving u, we require that dv/dn be constant 1 on a, and we 
recall (23-10). Since v * w + 1/r, this requirement is 


dw 5 1 

(23-12) 

— a* j. const. 

dn dn r 

To make w unique we require also that 


/ w da 0. 

Jr 

The Neumann function is 

N(P,Q ) s v = w + - 

(23-13) 

(23-14) 


r 


where v> satisfies (23-11) to (23-13). The solution u of (23-8) can evidently 
be expressed in the form 

u(P) * — f gvda * — f g(Q)N(P,Q) da . (23-15) 

4ir J * 4ir j <* 

When g is given in advance, it can be shown, conversely, that the function 
(23-15) satisfies (23-8). Hence, if we solve the particular Neumann 
problems involved in the construction of N(P } Q), we can solve the general 
Neumann problem for the region. 

Physically, the Neumann function represents the heat flow due to a source of strength 
4ir at P when the heat flows out at a uniform rate across the boundary. This shows 
that the condition 

dv 

— - 0 on <r (23-16) 

dn 

analogous to (23-3) cannot be required in general; when the region is bounded, (23-16) 
violates the principle of conservation of heat. For unbounded regions (23-16) is possible, 
as we see by considering the Neumann function for a half plane , 

N(P,Q) - - + — (23-17) 

r n 

It is left for the reader to verify that (23-17) satisfies (23-16) on the plane r 0 and 
to solve the Neumann problem. 


ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 

24. Classification and Uniqueness. If a, b , and c are real continuous 
functions of x and y, and if H is a continuous function of the indicated 
arguments, the partial differential equation 

Q&xx “f* 2 bz X y *4” (%yy 8=1 H (x^y fZjZxjZy} 

1 See remarks at the end of this section, 


(24-1) 



SBC. 24] BJXtPTIC, PAHABOLTC, AND HYPERBOLIC EQUATIONS 505 

includes many equations of mathematical physics. It is convenient to 
classify equations of the type (24-1) according to the sign of the discriminant, 
b 2 — ac f When b 2 — ac < 0, the equation is said to be elliptic; when 
b 2 — ac * 0, it is parabolic; and when b 2 — ac > 0, it is hyperbolic . This 
nomenclature is suggested by analogy with the conic 

ax 2 + 2 bxy + cy 2 - H ) a, b , c t H const, 

which is an ellipse, a parabola, or a hyperbola according to the sign of 
b 2 — ac. 

As typical illustrations, the reader can verify that 

V'XX 4“ Uyy ** 0, Ux* :=: tiy, *Uxx ** 0 (24-2) 

are elliptic, parabolic, and hyperbolic equations, respectively. The first 
of these is Laplace’s equation; the second 1 is the equation for heat flow; 
the third 2 describes the motion of waves on a string. The general equation 
(24-1) in the elliptic, parabolic, or hyperbolic case has much in common 
with the corresponding Eq. (24-2), and that is the reason why the classi- 
fication is important. We shall now discuss the conditions for unique 
determination of u. 

Case I. Elliptic Equation. Physical considerations suggest that a 
solution of Laplace’s equation is wholly determined by its boundary values. 
That is, if U\ and w 2 satisfy 

Uxx 4“ Uyy = 0 (24-3) 

in a bounded region r, and if u x = u 2 on the boundary, then u x « u 2 in 
r. A mathematical proof is readily given, assuming that the function 
u = — u 2 is continuous in r and on the boundary. 

Without loss of generality, let the region r lie between the lines x « 0 and x «* 1 
(so that cos x 0 in r). If v is defined by 

u « v cos x, 

a short calculation shows that (24-3) yields 

v xx 4* iw — 2v x tan x — v — 0. (24-4) 

Suppose that v > 0 at an interior point Po. Then v assumes a positive maximum at 
an interior point Pi (since f ® 0 on the boundary and v is continuous). At Pi we have 

V > 0, V X » 0, V XX <0, Vyy < 0, 

and hence (24*4) cannot hold. This contradiction shows that v < 0 throughout the 
region. Similarly, v > 0, and hence v m 0. It follows that u m 0, as was to be proved. 
The same method can be used in three dimensions; the only change is that (244) 

1 Let y «* aH in (9-5). 

* Let y m at in (24). 



PARTIAL DIFFERENTIAL EQUATIONS 


506 


[chap. 6 


has an extra term t> M . A decidedly less elementary proof (which applies also to the 
Neumann problem) was given in Sec. 21, Prob. 2. 

What we have actually shown is that if the harmonic function u satisfies 
u < 0 on the boundary, then the same inequality holds at interior points. 
Considering u — m instead of u yields the following significant result: 1 

Maximum Principle. Let u be harmonic in a bounded region r and let 
m be constant . If u < m throughout the boundary of r, then u < m through- 
out T. 

This theorem is true for the general equation 2 (24-1), provided (24-1) 
is elliptic and H > 0. 

Case II. Parabolic Equation . Let u be the temperature of a thin rod 
extending from x « 0 to x = l. With y * a 2 t the equation of heat conduc- 
tion is 

« m V) 0 < x <1, 0 < y < oo. (24-5) 

As typical initial and boundary conditions, we assume that 

u(x,0) ~f(x), u(0,y) « g(y), u(l t y) * h(y). (24-6) 

These conditions give the initial temperature and the temperature of the 
two ends. In the xy plane, (24-6) specifies the value 
of u on the boundary of a certain semi-infinite rec- 
tangle (Fig. 31). The physical interpretation sug- 
gests that u is thereby determined within the 
rectangle, and we shall now show that this is, in 
general, the case. 

Let u\ and u* satisfy (24-5) and (24-6). The function 

U » «i - Uj 

then satisfies (24-5) and (24-6) with / «* g *» h * 0. For 
simplicity, we shall suppose that u is continuous and bounded 
in the region of Fig, 31 and on its boundary, though these 
conditions could be weakened. 

If v is defined by u ® ve v , substitution in (24-5) yields 

v. (24-7) 

Suppose that v — Vo > 0 at some point Po of the rectangle in Fig. 31. We know that 
p * 0 on the three sides of this rectangle, and since u is bounded, the equation v ** e~ v u 
shows that v < vq if y is large enough. It follows that v assumes a positive maximum 
at some interior point Pi. At Pi, 

e*x < 0, *>0 

and hence (24-7) cannot hold. This contradiction shows that r <; 0, everywhere. 
Similarly, t> > 0, and hence v m 0, It follows that u\ m u*. 

*Cf. Sec. 21, Prob. 3. 

1 See: H. Bateman, “Partial Differential Equations of Mathematical Physics,” 
p. 135, Cambridge University Press, London, 1932. 




SBC. 25] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 507 

The method of proving this uniqueness theorem leads, just as in the foregoing discussion, 
to a Maximum Principle: If u < m on the boundary of the rectangle in Pig . 31, then 
u < m throughout the rectangle. A physical interpretation is readily given. 

PROBLEMS 

1. For what values of the constant k is u** -f ku^, +% « 0 elliptic? Parabolic? 
Hyperbolic? 

2. In what regions of the xy plane is 

(1 4* V)uxz *f 2xu xv -f (1 - y)Uyy - u z 

elliptic? Parabolic? Hyperbolic? 

8. Show that the solution of the elliptic equation 

u zx % *■ —ku 

is not always uniquely determined by the boundary values. Hint: Let the region be 
the square 0 x < *■, 0 < y < r, and separate variables. For a physical interpretation, 
see Sec. 14. 

4. A characteristic value for a region r is a constant X such that the problem 
Uxx + % + Usm + hu « 0 in r, u * Oon the boundary 

has a solution other than the trivial solution u « 0. Show that a characteristic value 
is always positive. Hint: If u ?£ 0, then u has a positive maximum or a negative mini- 
mum at some interior point P. 

6. The semi-infinite strip 0 < x < *, y > 0 has its edges kept at the constant tem- 
perature u «* 0, whereas its end y ** 0 is kept at the temperature u » sin x. In the 
steady state the temperature u satisfies u zx 4* Uyv * 0, and also 

u(0,y) « 0, u(x,y ) « 0, m(x,0) ® sin x . 

(a) By the method of separating variables obtain infinitely many distinct solutions to 

this problem. (6) Show that only one of these solutions satisfies lira u(x,y) 0 uni- 

v -* « 

formly in x. (r) If condition (6) is imposed, show that the problem has, in fact, only 
one solution. Hint: Use the maximum principle. 

26. Further Discussion of Uniqueness. Continuing the study of (24-1) 
we consider the hyperbolic equation 

Uxx Wyy * 0. (26-1) 

Since solutions of (25-1) do not satisfy the maximum principle, the fore- 
going methods cannot be used here. 

Case Ilia. Hyperbolic Equaiion t First Problem. Let the value and 
normal derivative of u in (25-1) be given on an interval (a, b) of the x 
axis (Fig. 32). Thus, 

w(x,0) «* /Or), u v (:r,0) * g(x), a < x <b. 

If G(x) * f g(s) d $, d'Alembert’s formula (3-15) yields an expression 

J a 

2u(x,y) = f{x + y) + G{x + y) + fix — y) — G{x - y), (25-2) 
which will now be used to discuss the uniqueness of u. 



SOB PABTIAL DIPPERENTIAL EQUATIONS [CHAP. 6 

By hypothesis, f(x) and G(x) are determined for a < x < b but not 
outside this interval. Hence 

f(x + y) 4* G(x + y) is determined for a < x + y < b, 

(25-3) 

f(x — y) — G{x — 2/) is determined for o < x — y < b, 
but not elsewhere. In the xy plane, the loci 

a <x + y <b, a <x~y <b 
represent two strips, bounded by the two pairs of lines 
x + y =* o, x + y ~ b and x — y ~ a, x — y « b (25-4) 

(Fig. 32). Both expressions (25-3), and hence u in (25-2), are uniquely 
determined in the intersection of these strips, but only there. This shows 
that the region of determinacy is the doubly shaded region in the figure. 1 



Similar behavior is found for the general hyperbolic equation, the role of the lines 
(254) being taken by the charactemtias introduced in Sec. 29. It is often possible to 
express u by an integral formula involving the initial value and normal derivative. 
The method requires construction of the Riemann function* which is in some respects 
analogous to the Green function of Sec. 23. 

Case Illfc. Hyperbolic Equation f Second Problem . The equation 
u x jc “ u vv has the general solution 

u(x,y) ** fi(x - y) + f 2 (x + y) (25-5) 

as was shown in Sec. 1. If u is given on two adjacent sides of the rectangle 
in Fig. 33, we shall use (25-5) to show that u is determined in the whole 
rectangle.* 

1 Cf. Theorem I, Sec. 4. 

. 1 See A. G. Webster, “Partial Differential Equations of Mathematical Physics/' 

p, 248, Teubner Verlagsgesellschaft, Leipzig, 1927. 

* Cf. Theorem II, 43ec. 4. 


BBC* 25] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 509 

Choose a point R in the rectangle, and draw RQ and RS parallel to the 
sides of the rectangle, as in the figure. With P the apex of the rectangle, 
x — y is constant on PQ and x — y is also constant on SR* Hence the 
same is true of fi(x — y) : 

fi(x — y) « a at P and Q , fi(x — y) * & at R and S. 

Similarly, 

f 2 (x + y) ** y at P and S, f 2 (x + y) » 5 at Q and J®. 

By using these values in (25-5) we can verify the identity 

u(R) ■» n(Q) + u(S) — u{P)* (25-0) 

This shows that u is determined by the data at every point in the rectangle 
and at no point outside the rectangle. 

By a procedure known as Picard's method , the problem just discussed 
can often be solved for the general hyperbolic equation (24-1). It is 
supposed that the equation has the form 1 

u xy = H(x,y,u,u Xf u v ) (25-7) 

and for simplicity we assume the homogeneous boundary conditions 

u(x, 0) - 0, u(0,y) = 0, 0 < x < a, 0 < y < b. (25-8) 

Thus, u = 0 on two adjacent sides of the rectangle in Fig. 34. The con- 
ditions (25-8) enable us to write (25-7) in 
the form 

u(x,y) =/ q j^H(x u y u u,u x ,u v ) dxidy x 

where the arguments of u, u x , and u v in the 
integral are x x and y x . 

Picard’s method consists of choosing a first 
approximation w <0) , evaluating the integral, 
and using the result as the second approxi- 
mation u (1 K A similar process yields w (2) , 
and so on. If u (n) is the nth approximation, 
the next approximation is 

u (n+n (x,y) = f rH(xi,yi,u (n) ,u< n) ,ul n) ) dx x dy t . (25-9) 

J0 J 0 

Subject to mild restrictions on H it can be shown 2 that the solution is 
1 Bee Sec, 29, Case III. 

I R. Oourant and D. Hilbert, “Methoden der mathematischen Physik,” vol. II, 
p. 317, I. Springer- Verlag OHG, Berlin, 1937. 




510 

given by 


PARTIAL DIFFERENTIAL EQUATIONS 


(CHAP. 0 


By (25-9), 


u(x,y) — lim u {n) (x,y). 

As an illustration, let the equation be 

«* 1 -f u. 

u (n+i) m r I* [i w (»)j dxi dyi xy + [* [* u (n) dxi dyi. 

/o </o /o /o 

Starting with u (0) «• 0, we get u (l) «■ :ry, « <2) » ary 4* <X|/) 2 /4 f 

" < " - xy +fif’ [* m + £«£] dzid„i-xv+ ( ^ i + i - 
and so on. Evidently, the process gives 

u(x,y) 

That this is a solution can be verified by actual substitution. 

26. The Associated Difference Equations. Let A be a positive number 
and u = u{x„y) a function of x and y. The difference operators A x and A v 
are defined by 

u(x + h,y) - u(x,y ) u(x, y + h) - u{x,y) 


(xy)* 

(30* 


^(xvT 

„_i(n0*' 


A z u 


(26-1) 


Passing to the limit as A — > 0, we get the partial derivatives; that is, 

lim A x u * w z , lim A v u ~ (26-2) 

— *■ o * — o 

when the limits exist. If the second differences are defined by 
u(x + ft, 2/) — 2u(x,y) + u(x - A, y) 




A 2 


Aftf/u 


u(x , 2/ + A) - 2u(x,v) + u(x, y A) 


it can be shown, in general, that 
lim A xx u ** 


A 2 


lim AyyU «* Uyy 
* — 0 


(26-3) 


(see Prob. 1). Hence the three difference equations 

AxxU “■ AyyU ® 0, A XX U 88 AyUy A XX U "f" AyyU ® 0 


as A — > 0 become the respective differential equations 

U X X Uyy 0, U X X ” Uy, 4" Uyy ** 0. 

The correspondence of difference equations and differential equations is important 
because there are numerical methods of solving the former which are especially adapted 



f 


SEC. 26] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 511 

to high-speed computers (cf. Chap. 9, Sec. 19). As h 0, the solution of the dif- 
ference equation generally tends to the solution of the corresponding differential equa- 
tion. This fact gives a means of numerical approximation which has been extensively 
exploited. Because of space limitations we shall consider merely the determinacy of 
the solutions, our objective being to clarify further the distinction among elliptic, 
parabolic, and hyperbolic equations. 

Case I. Elliptic Equation. Using (26-3) the reader can verify that 

A xx u + A yy u «=» 0 (26*4) 

can be written in the form 

u(x t y) * %[u(z + h, y) + u(x — h, y) + u(x, y + h) + u(z, y — h)]. 

(26-5) 

This equation gives a relation between the five values of u at the five 
neighboring lattice points 1 illustrated in Fig. 35; in fact, the value at the 




O' 


















\ 






r 

i c 

/ 

\ c. 

\ 





V 

7 \ 

/ 

7 t 

\ 





0 


V 





X 




r 





i " 


TT 






Fig. 36 


central lattice point is the arithmetic mean of the values at the four neighbors . 
The corresponding property for Laplace's equation is the average-value 
theorem given in the example of Sec. 21. 

To state the Dirichlet problem for the difference equation (26-4) we 
say that a point is interior to a region if its four neighbors are points of 
the region. A boundary point is a point for which at least one neighbor 
belongs to the region and at least one does not. For instance the points 
• in Fig. 36 are boundary points. In the Dirichlet problem a function u 
satisfying (26-4) is given at every boundary point, and it is required to 
find u at the interior points. We shall now establish both the existence 
and the uniqueness of the solution. 2 

Suppose, then, that u is known at every lattice point bounding a given 
region (Fig. 36). If we write down the equation (26-5) for each interior 

1 That is, points of the form ( mh t nh ) with integers m and n. 

* The region is assumed bounded, so that the number of interior points is a finite 
number n. 



512 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

point (%,y)> we obtain a system of n linear equations in n unknowns. It 
will be seen presently that the determinant of this system is not zero, and 
hence there is one, and only one, solution. On the other hand, if the values 
are not prescribed at every boundary point, there are always more un- 
knowns than equations, and the solution is not determined uniquely. These 
properties are analogous to those obtained previously for the Laplace 
equation. 

To show that the determinant is not zero, we shall analyze the special case in which 
the boundary values are zero. In this case the system of linear equations obtained by 
'writing (26*4) at each interior point is homogeneous. If the determinant is zero, the 
system will have a solution other than the trivial solution u m 0. Without loss of 
generality we can suppose that this nontrivial solution u is positive at some point. 

Let the maximum value of u over all the lattice points be denoted by m >0, and 
let P be a point where u ~ m. Evidently P cannot be on the boundary, since u > 0 
at P. Hence, P is interior, and the value of u at P is the average of the values at the 
four neighbors. Now u < m at these neighbors, since rn m the maximum. If u < m at 
any neighbor, then the average is <m t so that u(P) < m. This contradiction shows 
that u ® m at the four neighbors of P. 

We can now repeat the process, starting with one of these" four neighbors instead of P. 
Proceeding in this way we find that u — m at every lattice point. But that is impossible, 
since u » 0 3* m on the boundary. Hence the assumption that the determinant was 
zero led to a contradiction. 


PROBLEMS 

1. (a) Show that A xx u(z,y) « A x [ A x u(x — h , y) J. (b) If u has a Taylor series expan- 
sion about the point ( x,y ), show that A xx u u xx as h — ► 0. Hint: Use the first six 
terms of u(x + h, y k) « a -j- bh ■+■ rA -f ■ • ■ . 

2. Suppose AxxU + A w u *» 0, and suppose u is known for x = 0 and for x *= h 
(y w h, 2 h y 3/t, . , .). In what region of the ry plane is u determined? Hint * See Fig, 35. 

3. Let A xx u 4“ AyyU * 0, and suppose u(0 K v) * 1 for all y , u{2h,y) = 2 for all y, 
u(h t 0) - u(h,4h) - 0. Find u(h,2h). 

27. Further Discussion of Difference Equations. According to the fore- 
going discussion, the elliptic case leads to a set of simultaneous equations 
for determination of the unknown function u. In the parabolic and hy- 
perbolic cases, as we shall now see, the values of u can be obtained suc- 
cessively. 

Case II. Parabolic Equation. By (26-1) and (26-3) the equation 

A xx u a* A y u (27- J) 

takes the form 

u{x + h,y) ~ { 2 - h) u{x,y) + u(x ~~ h, y) « hu(x , y + h ). (27-2) 

This shows that if u is known at the three coliinear points in Fig. 37, then 
u can be found at the fourth point. 

By analogy with the problem of heat flow discussed in Sec. 24, let u 
be given at the points • on the boundary of the semi-infinite rectangle 



513 


SEC. 27] ELLIPTIC, PARABOLIC, ANO HYPERBOLIC EQUATIONS 

in Fig. 38. Referring to Fig. 37, we see that u can be found at the lattice 
points with y *■ h in the rectangle. Repetition gives u for y * 2 k, and 
so on. Thus, u is determined throughout the rectangle, just as in the case 
of Fig. 31. The process works equally well when the rod is infinite and 
u is given at all the lattice points on the x axis. 



- t 

L(*,y+ h) 

(*- 
- r - f 

V 

-A,y) 

f- 

7 

A*+Ky) 

t c 

h 

m 

■ 

) 






Fia. 37 


. 

y 


n 

— 

p-i 

— ' 

— 




r 






■ 

■ 

■ 

■ 

■ 

■ 



■ 

■ 

■ 

■ 

■ 

■ 








r 








■ 


h 

■ 

■ 

m 

m 

H 

i 


E2 

nr w w w i 



Fia. 38 


Because the pattern of Fig. 37 points upward, so to speak, it is impossible to proceed 
in the negative y direction when the rod is infinite. The very first step leads to a system 
of infinitely many equations in infinitely many unknowns. Inasmuch as y « ct 1 t > where 
t is time, this fact expresses the irreversibility of thermodynamic processes. 

Further insight into the one-directional character of i is given by (9-10) and '\8"4). 
In general, these expressions are infinitely differentiable for t > 0 but divergent for 
t < 0. This behavior of the heat equation contrasts to that of the wave equation. As 
we have repeatedly observed and will see again in the sequel, the latter is meaningful 
for negative t. 

Cask Ilia. Hyperbolic Equation, First Problem. Writing the equation 


A x ~u — A vy u = 0 


(27.3) 


in the form 

u(x + h,y) + u(x — h,y) — u(x y y + h) ~~ u(x y y ~ h) ~ 0 (27-4) 

we see that the corresponding pattern is that 
shown in Fig. 39. If u is given at any three 
of the four lattice points, then (27-4) gives u 
at the fourth point. Inasmuch as the pattern 
is symmetric, one can proceed in the positive 
y direction and in the negative y direction 
with equal ease. 

To discuss the analogue of the initial-value 
problem (Sec. 25, Case Ilia), let u and A y u be 
given in an interval of lattice points on the x 

axis. This is equivalent to specifying u itself on two adjacent rows of lattice 
points, as indicated by the black dots in Fig. 40. Considering Fig. 39 in 



t 

i 

Jx.y+h 

2 

u- 

\ 

-h t y) 

V _ ,... 

) 

/ 


1 

h 

/ 

t 

\ 

A*,y-h 



\ 

r r 


Fia. 39 


PARTIAL DIFFERENTIAL EQUATIONS 


514 


[chap. 6 


conjunction with Pig. 40, we see that the region of determination for u 
consists of the lattice points in the square. The analogy with Fig. 32 is 
evident. 

Case III&. Hyperbolic Equation , Second Problem . If u is given on two 
adjacent sides of a rectangle as shown in Fig. 41, we can apply Fig. 39, 



starting at P . It will be found that u is determined at the indicated 
points O, and at no others. This behavior corresponds to that found in 
Sec. 25, Case III6. 

PROBLEMS 

1. In Fig. 38 let u «■ 0 on the vertical rows of points •, and u « 1 on the horizontal 
row •, for 0 < x < l. Assuming h « 1 in (27-2) find u(h,Qh). 

a. In Fig. 40 let u * 1,0, 0,2, 0,0, 3, 0,0 on the bottom row of points • (in order), and 
let &yii ** 0 at these points. Find u(3/i,57t). 

3. Let u(P) « 0 in Fig. 41. Find the value of u at the opposite corner if u * 0 at 
the points • on the left of P, and u « 2 at the points • on the right of P. 

28, An Example: Flow of Electricity in a Cable. Many physical prob- 
lems lead to an equation that changes type, 
according to the values of the physical 
parameters. Since the character of the solu- 
tions undergoes a corresponding change, this 
phenomenon has great practical importance. 
As an illustration we shall consider the flow 
of electricity in linear conductors (such as 
telephone wires or submarine cables) in 
which the current may leak to ground. 

Let a long, imperfectly insulated cable (Fig. 42) carry an electric current 
whose source is at A. The current is assumed to flow to the receiving end 



Fig. 42 


SBC. 28] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 515 

at P through the load B and to return through the ground. It is assumed 
that the leaks occur along the entire length of the cable because of im- 
perfections in the insulating sheath. Let the distance, measured along 
the length of the cable, be denoted by x; then the emf V (volts) and the 
current I (amperes) are functions of x and t. The resistance of the cable 
will be denoted by R (ohms per mile), and the conductance from sheath 
to ground by G (mhos per mile). It is known that the cable acts as an 
electrostatic condenser, and the capacitance of the cable to ground per 
unit length is assumed to be C (farads per mile) ; the inductance per mile 
will be denoted by L (henrys per mile). 

Consider an element CD of the cable of length Ax. If the emf is V at 
C and V + AV at D, then the change in voltage across the element Ax 
is produced by the resistance and the inductance drops, so that one can 
write 

AV * — ^ IR Ax + — L A x^ * 

The negative sign signifies that the voltage is a decreasing function of 
x. Dividing through by Ax and passing to the limit as Ax — > 0 give the 
equation for the voltage: 

dV dl 

— - -IR - L — (28-1) 

dx dt 

The decrease in current, on the other hand, is due to the leakage and the 
action of the cable as a condenser. Hence, the drop in current A I across 
the element Ax of the cable is 

dV 

A I * - VG Ax C Ax, 

dt 

dl dV 

so that — * -FG - C (28-2) 

dx dt 

Equations (28-1) and (28-2) are simultaneous partial differential equa- 
tions for the voltage and current. The voltage V can be eliminated from 
these equations by differentiating (28-2) with respect to x to obtain 

/*, - -V X G - CV tx . 

Substituting for V x from (28-1) gives 

/** - IRQ + LGI t - CV tx 

from which V ix can be eliminated by using the expression for V xt obtained 
by differentiating (28-1). Thus one is led to 

I xx - JJClu * (LG + RC)I t + IRQ . 


(28-3) 



516 


PARTIAL DIFFERENTIAL EQUATIONS 


[chap. 6 

A similar calculation shows that (28*3) is also satisfied by the voltage V\ 
Evidently, (28-3) is hyperbolic when LC & 0 but parabolic when LC = 0. 

When the cable is lossless, R ~ G ~ 0. Equation (28-3) and the cor- 
responding equation for V are, then, 

I xx « LCI th V xx - LCV U . (28-4) 

Comparing the equation for wave motion (Sec. 1), we see that the cable 
propagates electromagnetic waves with velocity 

a « (LC)-X 

The hyperbolic equation (28-4) is appropriate if the frequency is high and 
the loss is low. 

For an audio-frequency submarine cable it is more appropriate to take 
G ss £ = 0. The equations are then parabolic: 

hx - RCI t , V xx = RCV t , (28-5) 

Instead of representing waves, the propagation of V and I is now identical 
with the flow of heat in rods. Comparing with (9-5) gives 

a - (RC)-X 

Example: Consider a submarine cable l miles in length, and let the voltage at the 
source A, under steady-state conditions, be 12 volts and at the receiving end R be 6 
volts. At a certain instant t m 0, the receiving end is grounded, so that its potential 
is reduced to aero, but the potential at the source is maintained at its constant value 
of 12 volts. Determine the current and voltage in the line subsequent to the ground- 
ing of the receiving end. 

It is required to find V in (28-5) subject to the boundary conditions 

V(0,t) « 12, V(l,t) - 0, t > 0. (28-6) 

The initial condition is 

F(x,0) - 12 - 6 - (28-7) 

since the steady-state solution of (28-5) is a linear function of x (Sec. 9, Example 1). 

The voltage V(x,t) subsequent to the grounding can be thought of as being made up 
of a steady-state 1 voltage Vs(x) and a transient voltage V r(z,0 which decreases rapidly 
with time. Thus, 

VO M) - V s (x) + Vr(x,t). (28-8) 

Since Vg(x) is linear, its value is given by the boundary conditions as 

V s (x) » 12 ~ 12 j- (28-9) 

Equations (28-6) and (28-7) now yield 

, W) - V T (l,Q - 0, Vrixfl) - j- 


1 Compare Sec, 9, Example 2. 



517 


SBC. 29] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 

Since Vr satisfies (28-5), we can use the solution of the heat equation (9-10) with o? m 
1/i RC, The result is 

Vrfat) m 2 ( 7 f 7 xi sin dxi) e“0/«c?K»ir//)*< gja 22, 

n»l \t JO l If l 

The function V is now given by (28-8). 

PROBLEMS 

1. By using (28-1) with L *® 0, find / in the Example. 

2. Find the emf in the cable whose length is 100 miles and whose characteristics are 
as follows: R ** 0.3 ohm per mile, C * 0.08 pi per mile, L * 0, G *> 0. If the voltage 
at the source is 6 volts and at the terminal end 2 volts, what is the voltage after the 
terminal end has been suddenly grounded? [Use (28-5). j 

3. Using (28-5), find the current in a cable 1,000 miles long, whose potential at the 
source, under steady-state conditions, is 1,200 volts and at the terminal end is 1,100 
volts. What is the current in the cable after the terminal end lias been sud- 
denly grounded? Use R *» 2 ohms per mile and C ** 3’10~ 7 farad per mile. 

29. Characteristics and Canonical Form. If a, b, c are continuous func- 
tions of x and y, with a ^ 0, then the partial differential equation 

au zx + 2 bu xy + cum - H{i,y,u,u x ,u v ) (29-1) 

can be simplified by use of the equation 

a dy 2 — 2b dy dx + c dx 2 = 0. (29-2) 

Setting dy == p dx in (29-2) and solving the resulting quadratic give 

p = a - " 1 ^ + (k 2 — ac)**] or p - a~ l [b — (6 2 — ac) H ], (29-3) 

Since p =» dy/dx , Eqs. (29-3) are ordinary differential equations of the 
first order, and hence the solutions may be expected to contain an arbitrary 
constant c. If the solutions are written in the form 

X (x,y) = c or Y(x,y) « c, (29-4) 

the resulting curves (29-4) are called the characteristics of (29-1). 

For example, when (29-1) is the wave equation 

a 2 Uxz — u tt « 0 (29-5) 

the differential equation (29-2) is 

o 2 dt 2 — dx 2 « 0. 

Since this reduces to dx/dt «* ±a, the characteristics are the straight lines 
x — al ** c t x + at c. 

It was shown in Sec. 1 that the change of variable 

r m % — at, $ * x 4* at, u(x,t) « U(r,s) 



518 PARTIAL DIFFERENTIAL EQUATIONS [CHAP. 6 

reduces (29-5) to the form U r § ** 0, and a physical interpretation of the characteristics 
was given in Sec. 4. 

Equation (29-1) is said to be in canonical form if it has one of the three 
forms 

Uxz “h %yy ^ Ujj * H, ttjy 388 H 

where H is a function of x, y , u, u x , and u y . It is a basic fact that the 
reduction to canonical form can be achieved by means of the characteristics, 
and we shall now describe 1 the procedure. 

Case I. Elliptic Equation. When b 2 — ac < 0, the two values of p 
in (29-3) are conjugate complex, and hence the same is true of X and Y 
in (29-4). That is, 

X « r{x,y) + is(x,y), Y » r(x,y) - is(x,y) 

where r and » are real. In this case the reduction can be achieved by choos- 
ing r and s as new independent variables. If u(x,y) - l/(r,s), Eq. (29-1) 
gives an equation for U in which the second derivatives occur as U rr + 
U 

Case II. Parabolic Equation. When b 2 — ac = 0 the two values of 
p in (29-3) are real and equal. Hence the same is true of X and Y in 
(29-4). In this case the reduction can be achieved by the change of vari- 
able 

r ** X{x,y), s « any function independent of X . 

The second derivatives of U now occur as U et . 

Case III. Hyperbolic Equation. When b 2 — ac > 0, the roots (29-3) 
are real and unequal, and the same is true of X and F. The reduction is 
achieved by taking 

r « X(x,y), s « Y(x,y) 

as new independent variables. The second derivatives of U occur only 
as U r9 . 

To illustrate the procedure we shall consider the equation 

Uxx *“*’ kxilgy *f 4 X^Uyy ** 0 (29**6) 

when k — 0, 4, or 5. According to (29-3), 

p » -j^kx ± - 4x*) H . (29-7) 

When k — 0, this gives p « ±2 ix. The equations y‘ - db2 ix have the solutions 
y — ix* » c, y +ix* ** c 

1 A proof may be found in A. G. Webster, “Paitial Differential Equations of Mathe- 
matical Physics/ 1 p. J242, Teubner Verlagsgesellschaft, Leipzig, 1927. 



519 


SBC. 80] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 
where c is constant. Taking real and imaginary parte, 

r « V, * - **. 

With u(x,y) m U(r t e) t the derivatives are 

« 4x 2 f/g, -f 2U 9f Uxy « 2 xUtr, Uw - Urr 

and substitution into (29-6) with k -* 0 gives the canonical form 

Urr + U» - ~(2x*)~ l U 9 m ~(2 

When k » 4, the two roots (29-7) are both p * — 2x. Solving this differential equa- 
tion we see that (29-4) is 

y + i 2 * c, j/+r 2 »c. 

Since y -f a; 2 and y are independent, we can take 

r « y *f **, * « y. 

It is left for the reader to show that the canonical form is 


U» « — (2x 2 )“" 1 < r / r - — (2r - 2s)~ 1 £/ r . 


Finally, the case & « 5 leads to two distinct real roots p 
p «* dy/dx and solving, 

y + “ c, y + 2z 2 ™ c. 


The change of variable 

+ M* 2 , s * y + 2x 8 

now leads to the canonical form 

U„ * (68 ~ fir)~\U r + 4C/.). 


-*» P 


(29-8) 
— 4x. Setting 


(29-9; 


PROBLEMS 


1. Derive (29-8) and (29-9). 

2. Describe the behavior of the characteristics of (28-3) as LC varies from asero to 
infinity. 

8. Reduce to canonical form 

3tt»y « U M -F 2uyy t 2u xy » U tX + Uyy, 2Ugy - U** + 2^. 

30. Characteristics and Discontinuities. The function u * f(x — a<) 
represents a wave propagating in the positive x direction with velocity a. 
If f{x) has the form shown 1 in Fig. 43, the motion exhibits a wave front 
(Fig. 44), whose locus can be found by setting the argument off equal to c; 

x — at «* c. 

In the xt plane, we recognize that this equation describes a characteristic 
of the wave equation (29-5). 

» The intent is f(x) - 0 for x > c, /'(c) - 0, /"(c+) - /"(«-) * 0. 



PARTIAL DIFFERENTIAL EQUATIONS 


520 


(chap. 0 


Discontinuities of the type just considered arise in many investigations, 
ranging from the theory of the cracking of glass to the theory of super- 
sonic flight. The locus of the discontinuity is always a characteristic, as 
we shall presently see, and hence, the foregoing example is typical of the 
general case. 



c * 

Fia. 43 



Since the locus is a characteristic, a discontinuity of the type in question 
may arise on two families of curves for hyperbolic equations, it may arise 
on one family for parabolic equations, and it cannot arise for elliptic equa- 
tions. For example, the equations of fluid flow are elliptic at velocities 
less than the velocity of sound in the fluid. But at velocities exceeding 
the velocity of sound the equations become hyperbolic, and the fact that 
a discontinuity is now possible permits the formation of a shock wave. 

To discuss these questions mathematically, consider a solution surface 
u «= u{x t y) satisfying (29-1). We suppose that u is continuous and has 
continuous first derivatives but has a discontinuity in one of the second 
derivatives on a certain curve C. This single solution is to be regarded as 
two solutions u 2 (^,y) which are tangent along the curve C but 

do not have equal second derivatives along C. We let the surfaces de- 
fined by U{ and u 2 extend past C, so that their first derivatives are well 
defined on C. 

The symbol ( ) denotes the jump of the function in the parentheses; 
that is, 

(w) = u\ ~ u 2 evaluated on C. (30-1) 


Differentiating (30-1) with respect to x gives 

(u x ) ~ U U - u 2x * (u) x (30-2) 

and similarly for other derivatives. Our hypothesis is that 


(w) — ( U x ) *= ( Uy ) — 0 


(30-3) 


but that one of the quantities ( u xx ), (u zy ) t or (u yy ) is not zero. 

Differentiating the relation ( u x ) »* 0 with respect to x by the chain 
rule yields 

(uxx) + (u xy )y' - 0 (30-4) 

when we recall that y is a function of x on C and take due account of 
(30-2). Similarly, differentiating {u y ) = 0 with respect to x yields 


(uyx) + (u yy )y' « 0. 


(30-5) 



521 


SEC. 30] ELLIPTIC, PARABOLIC, AND HYPERBOLIC EQUATIONS 

Taking the ( ) of the partial differential equation (29*1), we get 

a{u x , r) + 2 b(u xy ) + c{uyy) * (if) * 0 (30-6) 

when if is continuous. The fact that (H) « 0 follows from (30-3) if we 
recall that H does not involve the higher derivatives of u. 

Equations (30-4), (30-5), and (30-0) are three linear homogeneous 
equations in the three unknowns (u xx ), (u* v ) = (u yz ) , and (u vv ). By 
hypothesis not all these unknowns are zero, and hence, the coefficient 
determinant must vanish: 


1 y' 0 
0 1 y' 

a 2b c 


0 . 


(30-7) 


Expansion of this determinant yields the characteristic equation (29-2), 
so that C must be a characteristic curve. Conversely, if C is a character- 
istic, the determinant (30-7) is zero, the related homogeneous equations 
have a nontrivial solution, and a discontinuity is possible. 

Example. Fundamental Solutions. A fundamental solution of a partial differential 
equation is a solution of the form /[X(x,y)] f when X is a fixed function and /an arbitrary 
function. For example, the equation u zx « Uyy has the fundamental solutions 

/i(z - y) and / 2 (x + y) t30-8) 

in which X « x — y and X =* x + y, respectively. We shall now see that if (29-1) has 
the fundamental solution f[\(x,y)], then the curves 

X(x,y) » c, c const (30-9) 

are characteristics. For proof, it suffices to choose the arbitrary function / so that f"{x) 
is continuous except at i * f. Then the function u = /[X(x,j/)] has a discontinuity of 
the type previously considered on the locus X(x,?/) = c, and the desired result follows. 

This explains why the techniques used in Secs. I to 6 to study the wave equation are 
not applicable to Laplace’s equation. Namely, d’Alembert’s method is based on the 
fundamental solutions (30-8) and the Laplace equation, lieing elliptic, has no such 
solutions. 




CHAPTER 7 


COMPLEX VARIABLE 




f 


Analytic Aspects 

1. Complex Numbers 527 

2. Functions of a Complex Variable 534 

3. Elementary Complex Functions 535 

4. Analytic Functions of a Complex Variable 540 

5. Integration of Complex Functions. Cauchy’s Integral Theorem 545 

6. Cauchy’s Integral Theorem for Multiply Connected Regions 548 

7. The Fundamental Theorem of Integral Calculus 550 

8. Cauchy’s Integral Formula 555 

9. Harmonic Functions 559 

10. Taylor’s Series 501 

11. Laurent’s Expansion 555 

12. Singular Points. Residues 509 

13. Residue Theorem 573 

14. Behavior of f(z) at Poles and Essential Singular Points 574 

Geometric Aspects 

15. Geometric Representation 575 

16. Functions w — z n and z — y/w 577 

17. The Functions w = c * and z = log w 581 

18. Conformal Maps 583 

Applications 

19. Steady Flow of Ideal Fluids 587 

20. The Method of Conjugate Functions 588 

21. The Problem of Diriehlet 595 

22. Evaluation of Real Integrals by the Residue Theorem 599 


525 




f 


This chapter contains a concise presentation of the rudiments of com- 
plex-variable theory with an indication of its many uses in the solution of 
important problems of physics and engineering. This theory, with roots 
in potential theory and hydrodynamics, is among the most fertile and 
beautiful of mathematical creations. Its unfolding left a deep imprint on 
the whole of mathematics and on several branches of mathematical physics. 
To an applied mathematician this theory is a veritable mine of effective 
tools for the solution of important problems in heat conduction, elasticity, 
hydrodynamics, and the flow of electric currents. 


ANALYTIC ASPECTS 

1. Complex Numbers. The analysis in the preceding chapters was 
concerned principally with functions of real variables, that is, such vari- 
ables as can be represented graphically by points on a number axis, say 
the x axis of the cartesian coordinate system. The reader is familiar with 
the tact that calculation of the zeros of the function f(x) = ax 2 4 bx 4 c, 
when the discriminant b 2 — 4oc is negative, necessitates the introduction 
of complex numbers of the form u 4 iv, 
where u and v are real numbers and i is a 
number such that i 2 = — 1. 

A number of the form u + iv can be repre- 
sented by a point in a plane referred to a pair 
of orthogonal x and y axes if it is agreed that 
the number u represents the abscissa and v 
the ordinate of the point (Fig. 1). No con- 
fusion is likely to arise if the point ( u,v ), asso- 
ciated with the number u 4 iv, is labeled 
simply u + iv. It is clear that the point (u,v) can be located by the 
terminus of a vector z whose origin is at the origin 0 of the coordinate 
system. In this manner a one-to-one correspondence is established 
between the totality of vectors in the xy plane and the complex numbers. 

527 




m 


COMPLEX VARIABLE 


[CHAP. 7 

The vector z may be thought to represent the resultant of two vectors, 
one of which is of magnitude u and directed along the x axis and the other 
of magnitude v and directed along the y axis. Thus, 

z * u + iv, 

where u is spoken of as the real part of the complex number z and v as 
the imaginary part. Therefore, if the points of the plane are referred to 
a pair of coordinate axes, one can establish a correspondence between the 
pair of real numbers ( u,v ) and a single complex number u + iv , In this 
case the xy plane is called the plane of a complex variable, the x axis is 
called the real axis , and the y axis is called the imaginary axis . 

If v vanishes, then 

z « u + 0-t *= u 

is a number corresponding to some point on the real axis. Accordingly, 
this mode of representation of complex numbers (due to Gauss and Argand) 
includes as a special case the usual way of representing real numbers on the 
number axis. 

The equality of two complex numbers, 

a + ib = c + id, 

is interpreted to be equivalent to the two equations 
a = c and b = d. 

In particular, a + ib » 0 is true if, and only if, a « 0 and b « 0. 

If the polar coordinates of the point ( u,v ) (Fig. 1) are (r,0), then 

u «= r cos 0 and v =» r sin 0 

so that 

r » Vm 2 + v 2 and 6 * tan"” 1 -• 

u 

The number r is called the modulus , or absolute value , and 6 is called the 
argument, or phase angle , of the complex number z * u + iv. It is clear 
that the argument of a complex number is not unique, and if one writes 
it as 6 + 2 kr, where 0 < 6 < 2r and k = 0, =4=1, db2, . . ., then 6 is called 
the principal argument of z. The modulus of the complex number z is 
frequently denoted by using absolute-value signs, so that 

r = | z | « | u + iv | * Vy 2 + v 2 , 

and the argument $ is denoted by the symbol 

6 * arg z . 

The student is assumed to be familiar with the fundamental algebraic 
operations on complex numbers, and these will not be entered upon in 



t 

BBC. ij ANAXmC ASPECTS 529 

detail here. It should be recalled that (cf . Chap. 2, Sec. 15) 

zi -f *2 =■ (xi + iyi) + (x s + iy») = (*i + x 2 ) + i(.Vi + Va), 

Z\'Z 2 (*1 + m )( x 2 + rn) = (Xix 2 - y x y 2 ) 4- i{x x ya + x 2 y x ), 

zi x, + iy x xix a + y x y 2 . x 2 y t - x x y 2 

— as — ■ ass — — - »■ ■■■■■» ■ «*— i 

z 2 X 2 + iy 2 x\ + vi *2 + 2/2 

provided that \z 2 \ * v erf + y\ 0. 

On representing complex numbers z x and z 2 by vectors, we can see at 
once from Fig. 2 that they obey the familiar “parallelogram law of addition” 
formulated in Chap. 4. 



From elementary geometric considerations we deduce that 

I z l + z 2 | < | Z\ I + | Z 2 I ; (1-1) 

that is, the modulus of the sum of two complex numbers is less than or equal 
to the sum of the moduli. This follows at once from Fig. 2 on recalling that 
the sum of two sides of a triangle is not less than the third side. 

Also, 

k + z2l>M~M; d-2) 

that is, the modulus of the sum is greater than or equal to the difference of the 
moduli. This follows from the fact that the length of one side of a triangle 
is not less than the difference of two other sides. 

Equations (1-1) and (1-2) yield a useful inequality, 

|zi|-|*2i<|*i -*2|<|*i| + l*a|, (1-3) 

indicated in Fig. 3. 

When calculations are carried out with complex numbers, the notion of 
the conjugate complex number is useful. We define the conjugate 2 of the 
number z *= x + iy by the formula 

2 *» x — iy. 



530 COMPLEX VARIABLE (CHAP. 7 

The application of the rules for addition, multiplication, and division 
of complex numbers yields the following theorems: 

(' a ) «i + 2 2 « li + l 2y (1-4) 

or, in words, the conjugate of the sum of two complex numbers is equal to the 
mm of the conjugates; 

(*>) = hh, (1-6) 

that is, the conjugate of the product is equal to the product of the conjugates; 



or the con jugate of the quotient is equal to the quotient of the conjugates . 

We note that if l = z, then z is real. 

The geometric interpretation of multiplication and division of complex 
numbers follows readily from polar representation of complex numbers. 
Thus, 

z x z 2 «* ri(cos 0i + ism 0i)r 2 (cos 0 2 + % sin 0 2 ) 

=* rir 2 [cos (i 0\ + 0 2 ) + i sin (0i + 02)1- (1-7) 


That is, the modulus of the product is equal to the product of the moduli and 
the argument of the product is equal to the sum of the arguments . 

Also, 


Z 1 


ri(cos0i + i sin 00 
r 2 ( cos 0 2 + i sin 0 2 ) 


r x 

= — [cos (0i — 0 2 ) + i sin (0i — 0 2 )], 
r 2 


(1-8) 


as follows on multiplying the numerator and denominator in (1-8) by cos 0 2 
— i sin 0 2 . Thus, the modulus of the quotient is the quotient of the moduli 
and the argument of the quotient is obtained by subtracting the argument of the 
denominator from that of the numerator. 

On extending formula (1-7) to the product of n complex numbers 


we get 


** « r*(cos 0k + i sin 0*), k - 1, 2, . . . , n, 


* r t r 2 . . .r n [cos (0 X + 0 2 H b 0 n ) + i sin (0 X + $ 2 H b 0n)] 

and, in particular, if all za are equal, 

z n « [r(cos 0 + i sin 0)] n « r n (cos n$ + i sin n0). (1-9) 

Formula (1-9) is known as the de Moivre formula, and we have shown that 
it is valid for any positive integer n. We can show that it is also valid 
for negative and fractional values of n. 



631 


SBC. 1] ANALYTIC ASPECTS 

Indeed, from (1-8) we deduce that 
1 cos 0 -f i sin 0 1 

- * — : * ~ [cos (-$) + i sin (-6)], 

z r(cos 0 + i sm 0) r 

and since (1-9) is known to hold for positive integers n, 

a z~ n » [cos (— 0) + i sin (—0)]* 

« r^[cos ( — n$) 4* tsin ( — n0)]. 


This establishes the result (1-9) for negative integers n. 

To prove the validity of (1-9) for fractional values of n, it suffices to 
show that it holds when the integer n is replaced by 1/n, for on raising 
the result to an integral power m, we obtain the desired formula for frac- 
tional exponents. 

Let 

w m z lln m tyz 9 

so that w is a solution of equation 

w n ® z . 

On introducing polar representations, 

w = R(cos (fi + i sin p), 
z « r(cos 6 + i sin 0), 


( 1 - 10 ) 

(Ml) 

( 1 - 12 ) 


where 0 is the principal argument of z, we can write (1-11) with the aid 
of (1-9) as 

w n «* R n (cos n<p + i sin n<p) * r(cos 0 + i sin 0 ). 

We conclude from this that 


R n a* r, n<p =» 0 dfc 2&7r, A; *= 0, 1, 2, . . ♦ , 


and thus 
Hence, from (1-12), 


n/“ 0 d: 2 /ct 

J { * vr, ^ — » fc *c 0, 1, 2, • . .. 






( 


n 

0 dz 2 fcir 


r 1 cos ■ 


+ i sin ■ 


0 ± 2kr 


: 2kr\ 

—)• 


and on recalling (1-10), we see that 

, , , , . , / e ± 2fcir . e ± 2kx\ 

2 i/» „ [ r ( C08 e + i sin 0)] 1/n ■> r l/n ( cos (- i sm }• 

\ n n / 


(1-13) 



COMPLEX VARIABLE 


[CHAP. 7 

Since cob (6 ± 2kr)/n and sin (9 =b 2kr)/n have the same values for 
two integers k differing by a multiple of n, the formula (1-13) yields just 
n distinct values for \^z, namely, 


Vi 


r i/» 



8 + 2kr 

n 


+ tsin 


6 + 2 kr\ 


k «■ 0, 1, 2, . . n — 1. 

(1-14) 


The validity of formula (1-9) for fractional values of n follows directly 
from (1-14) upon raising z lln to an integral power m. 

We illustrate the use of formula (1-14) by two examples. 

Example 1. Compute In this case z * 1 and its principal argument 0*0. 
Formula (1-14) then yields 

nr~ 2 kx . , 2kx 

V 1 « cos h i sm > k » 0, 1, ...,« — 1. 

n n 


If we plot these « roots of unity, we see that they coincide with the vertices of a regular 
polygon of n sides inscribed in the unit circle, with one vertex of the polygon at. z *• 1. 
Figure 4 shows this for n *» 6. 

Example 2. Find all roots of ' + Since 1 + i * y/2 [cos (t/4) + % sin (t/ 4)], 
formula (1-14) gives 


»/• «/~ / (tt/4) 4* 21ct , , (r/4) -f~ 2kx\ 

Vi + f * V2 ^cos - j - 1 sm » k «■ 0, 1, 2. 


Thus the desired roots are 

wi « ^5(cos K 2 * -f i sin H 2 *-), 
1 Z >2 « V^(co8 -f* i sin K*), 
wz - -^(cos 4* t sin 1 K2»)- 
These roots are represented in Fig. 5. 



I?ia. 4 


Fig. 5 



ANALYTIC ASPECTS 


SBC. 1] 

The reader unskilled in simple calculations involving complex numbers 
is urged to work out the representative problems in the following list before 
proceeding to the next section. The symbols Re (z) and Im (z) used in 
some problems in this list denote, respectively, the real and imaginary 
parts of a complex number z. 


PROBLEMS 


1. Find the moduli and principal arguments of the following numbers, and represent 
the numbers graphically: 


(a) 1 + tV 3, (b) 2 + 2i, <c) -2, (d) i\ (e) 


1 


if) 


1 + i 


w a - i) 4 . 

2. Write the following complex numbers in the form a + bi: 


1 + i' 1 -V (t- V5) 8 ' 


is) 


(a) (1 - V3 0* (6) 


(1 + i) t 


(c) ■ 


V 3 * 


1 - i ' v 1 4- i 

3. Find the cubes of the following numbers: 

(a) 1, W Vii-\ + »V 3), (c) Mi-\ - t\/3). 

4 . Find the cube roots of t, and represent them graphically. 

6. Find all solutions of the equation z 4 4~ 1 *0. 

6. Verify that z 2 — 2z 4* 2 » 0 has the roots z « 1 ±t. 

7 . Compute and represent graphically the following numbers: 

(a) </l, (b) W, (r) J/i, (d) 

8. Find all the fifth roots of 1 -f- i, and represent them graphically. 

9. Use de Moivre’s formula |r(cos0 + 7 sin 0)| n = r n (cos nO -f i sin nd ) to o K oain 
cos 20 « cos 2 0 — sin 2 0 and sin 26 « 2 sin 0 cos 0. 

10. Write the following numbers in the form a - \~bi : 

(a) Vi, (6) Vi - i, (c) ■ 1 


Vl + * ' 

11. Prove that ( a ) z x 4- z 2 ** 22 4* *1, (6) Z1Z2 ® z&\ } (c) z x (z% 4* 23) 

12 . Show that if ziZ2 *= 0, then t\ * 0 or z 2 * 0, 

13 . Prove formulas (1-4), (1-6), and (1-6). 

14 . Find |z|, Z, Re (z), and Im (z) for the following: 

(a) z - 1 — 2 f; (6) * - 3 + 4i; (c) r - h — . 

3 + 4t 

16. Show that (a) tz » — iz, (6) |?| ® |z| 3 . 

16 . What is the locus of points for w^hich 
(a) j z | - 1? (b) \z | < IV (e)|#|>17 

flint : \z\ «“ y/ a* 2 + y 2 . 

17 . If z » x 4“ iy t what is the locus of points for which 
(a) Re (z) > 1? (6) Im (z) > l? (c) Re (z 2 ) « 1? 

18 . If z » x 4“ W, describe the loci: 

1 I z — 1 I 

(a) \z — 1 1 -* 2; (6) — * const; (c) — — - const, 

|Z| I 3 4* 1 I 

19 . Under what conditions does one have the relation 
(a) \z\ -h h\ m |*il 4"|*al? (&) |*i 4- h \ «* !*i| - 

20. If * « x 4* UA write the following in the form u 4* tv: 

<«} ^ ** +* - 1. w -Lj. 

Z i — Z Z T’ l 


*1*5 + 



534 


COMPLEX VARIABLE 


[CHAP. 7 

2. Functions of a Complex Variable. A complex quantity z -* x + iy 
in which x and y are real variables is called a complex variable . We shall 
speak of the plane in which the variable z is represented as the z plane. 
If in some region of this plane for each z ~ x + iy one or more complex 
numbers it? «* u + tv are determined, we say that tt? is a function of z 
and write 

u? » u + it? « f(z). 

Thus, w « x 2 — y 2 + i2 xy « (x + iy) 2 « z 2 

is a function of z defined throughout the z plane. Also, 
u? « u -f iv ** x — iy « 2 


is a function of z. In fact, every expression of the form w(x,y) + iv(x,y) 
in which u and t> are real functions of x and y is a function of z, since x as 
J^(z + 2) and y as (1/2 i)(z — z) are functions of z. 

A complex function w = f(z) is single-valued if for each z in a given region 
of the z plane there is determined only one value of w. If more than one 
value of w corresponds to 2 , the function w = f(z) is multiple-valued. Thus 

w ~ z 2 ~ y 2 + i2xy = z 2 
and w ~ x 2 — y 2 — i2xy « g 2 

are single-valued functions of z. The function w = y/z for each z^O 
determines two complex numbers, for on setting z = r( cos 0 + i sin 6) 
and recalling formula (1-14), we get 


so that 


It? SC 

W\ ** 

u?2 “ 


(■ 


0 + 2fcjr 0 + 

cos f- i sin — 


)■ 


( 9 . . 9 \ 

I cos - + 1 sin - I < 

V 2 2/ 

jj/os (~ + -f t sin (HI- 


* - 0 , 1 , 


Thus w = y/z is not single-valued. 

The functions in the foregoing examples are defined throughout the 
z plane. The function w =* l/z is not defined at the origin z » 0, while 
10 « l/( J 2 j — 1) is not defined when \z\ « 1, that is, when the points 2 
lie on the circle of radius 1 with center at the origin. 

Of course, it? = f(z) may be defined by different formulas in different 
regions of the plane, or it may not be defined at all in certain regions. 

In dealing with regions of the z plane we shall distinguish interior points 
from those that lie on the boundaries of the region. A characteristic prop- 
erty of the interior points is that about each interior point P one can draw 
a circle with center at P and with nonzero radius r so small that the circle 



ANALYTIC ASPECTS 


SBC. 3] 


535 


contains only those points that belong to the region. The points on the 
boundary of the region are not interior because every circle with the 
boundary point as its center includes points that do not belong to the 
region. 

A region consisting only of interior points is said to be open . An ex- 
ample of such a region is the circular region whose points z satisfy the 
condition \z\< R. When the boundary of the region is included in the 
region, the region is called dosed. An example of a closed region is the 
region consisting of points z such that | z | < R. 

If every point of the region is at a finite distance from the origin, the 
region is said to be finite or bounded . Thus all points of the bounded region 
lie within a circle |z| - R if the radius R is chosen sufficiently large. The 
region consisting of all points in the z plane is unbounded , and so is the 
region consisting of the points satisfying the condition \z \ > 1. 

A plane region is simply connected if every closed curve drawn in the 
region encloses only points of the region. Thus, a region bounded by an 
ellipse is a simply connected region, while a region bounded by a pair of 
concentric circles is not simply connected. A region that is not simply 
connected is called multiply connected . 


PROBLEMS 


1. Express the following functions in the form u(x,y) + iv(x,y ): 

(a) Z* - * + 1, (f>) (r) (d) 0) ,(/)* + M(z - i), ig ) -i, 

Z ~t l 


(h) 


z 2 — 2z + 1 


1 


z + i 


2 

«** 


z -f* 2 

2. Describe the regions in the z plane defined by the following condition*: 
(a) Re (z) <3; (6) Im(z)>l;(c) \z\>l](d)\ <|z|<2;« |*-1|<1;</) \z ~ 

< 1 ; (fir) |* +i|> 2. 

3. Elementary Complex Functions. In Sec. 1 we defined the operations 
of addition, multiplication, division, and root extraction for complex 
numbers. These suffice to determine, for any z, values of such algebraic 
expressions as 

Oo z m + Oi2 m_1 H ho 


w ~ 


boz n + biz n 1 . . . -(- b 


in which the powers m and n may be integers or fractions. However, they 
do not provide direct means for defining the complex counterparts of the 
real elementary transcendental 1 functions e x , sin x, log x, tan~ l x, etc. 

1 A variable w satisfying the equation P(z y w) ** 0, where P is a polynomial in z and 
u\ is called an algebraic function of z. A function that is not algebraic is called transcend 
dental. The trigonometric and logarithmic functions and their inverses are called ele- 
mentary transcendental function #. 



COMPLEX VARIABLE 


[CHAP. 7 

A useful definition of a complex function such as e *, for example, must 
specialise to e* when z assumes real values. Also, it is desirable to pre- 
serve the familiar law of exponents e**e** * e*i+\ 

A definitive formula for e* that fulfills these criteria is 


e* » e* +iv •» e*(cos y + i sin y). (3-1) 

Moreover, as we shall presently see, it suggests sensible definitions for 
all the other elementary transcendental functions. We note first that for 
*»0 the definition (3-1) yields 

e %v « cos y 4* i sin y. (3-2) 

On replacing y by ~y we get 


e~~ tv * cos y — i sin y . 

Adding and subtracting (3-2) and (3-3) we get the Euler formulas 
cos y « Yt{e iv + e m ~ iy ) f 

1 

sin y » — ; (e %v — e Hf ). 


(3-3) 


(3-4) 


These formulas suggest that we define the trigonometric functions of z 
as follows: 


cos z « - (e'* + e '*), 
2 


sm 2 




sin 2 

tan 2 « » 

cos 2 


1 1 1 

cot 2 *» sec 2 a • CSC z ** 

tan 2 cos 2 sin 2 


(3-5) 


Using these definitions it is easy to check 1 that all the familiar formulas 
of analytic trigonometry remain valid when real arguments are replaced 
by the complex ones. For example, 

sin 2 z + cos 2 2 = 1, 


sin (21 + z%) * sin zi cos z% + cos t\ sin z 2f 

and so on. 

The logarithm of a complex number 2 is defined in the same way as in 
real variable analysis. Thus, 


means that 


w ** log 2 
2 « e** y 


(3-6) 

(3-7) 


* See Prob. 1 at the end of this section. Also, of. alternative definitions of **, «n s* 

and cos * in Sec. 17, Chap. 2. 



BEC. 3] ANALYTIC ASPECTS 537 

where e is the base of natural logarithms. Setting w «* u + w in (3*7) 
gives 

z » e w+ ^ « e*(cos + t sin v) (3*8) 

by (3-1). On the other hand, we can write z as 

z « x + iy *» r(cos 0 + t sin 0), 

so that (3-8) gives 

r(cos 0 + i sin 0) « e“(cos i; + t sin v). 

It follows from this that 

e u « r, v * 0 + 2fcir, fc * 0, dbl, db2, . . .. (3-9) 

Since u and v are real, we conclude from (3-9) that u * Log r, where the 
symbol Log is used to denote the logarithm encountered in real-variable 
theory. We can thus write (3-6) in the form 



w * u + iv ®* log z *= Log r + (0 + 2kr)i 

(3-10) 

or 

1 y 
log z = - Log (x 2 + y 2 ) + i tan"” 1 -> 

2 x 

(3-11) 


since r = Vj 2 + y 2 and 0 + 2kir = tan"" 1 (yfx). 

Thus log 2 has infinitely many values corresponding to the different 
choices of the arguments 0 of z. Setting k ~ 0 in (3-10) and assuming 
that 0 < 6 < 2w, we get a single- valued function 


log z * Log r + 6i f 0 < 6 < 2ir, 


which is called the principal value of log z. If z is real and positive, the 
principal value of log z equals Log r. 

The definition (3-10) serves to define complex and irrational powers c 
of the variable z by the formula 

z e * e c log * (3-12) 


which is equivalent to the statement that log z c ~ c log z. Inasmuch as 
log z is infinitely-many- valued, it follows that z c ) in general, is an infinitely- 
many- valued function. 1 The hyperbolic functions of z are defined by the 
formulas 

1 1 _ sinh z 

sinh z = - (e* — e *), cosh z * - (e f + e *), tanh z » — — » 

2 2 cosh z 


sechz 


1 


coshz 


csch z 


1 


sinh z 


(3-13) 


1 Note, however, that t e is single- valued when c is an integer. 



COMPLEX VARIABLE 


[CHAP. 7 

These functions are clearly single-valued. The inverse trigonometric and 
inverse hyperbolic functions are defined in the same way as in real-vari- 
able analysis, and they are multiple- valued . 1 

Example 1. Compute e l ~'*. 

On setting * - 1 and y * —1 in the formula (3-1) we get 
a 1 -* » e[cos ( — 1) +ia in ( — 1)) 

— e(cos 1 — i sin 1). 

Since cos 1 - 0.54030, sin 1 « 0.84147, and e - 2.718, 

e w - 2.718(0.5403 ~ i’0.8415) 

- 1.469 - t‘2.287 

to three decimal places. 

Example 2. Compute sin (1 — i). 

Since 

sin z ~ ~ ( e” - e~“) 

and * *» 1 — i, we have 

mn(l — i) - i (« <+1 - «— *) 

£X 


2i 

e — t 


1 -f % sin 1) — e~ l [cos ( — 1) 4- i sin ( — 1)1 1 
e 4* e“ l 


2 i 


- cos 1 -f - 


2 


- sin 1. 


We can obtain the same result by making use of the addition formulas of trigonometry. 
Thus, 


sin (1 — i) **» sin 1 cos (*~t) 4- cos 1 sin (— i) 


But by (3-5) 


» sin 1 cos i — cos 1 sin i. 

cost ~ i (e"* 1 4- c 1 ), sin i « ~~ (e~ l — e l ). 
2 2 % 


Substitution In the foregoing formula yields the result obtained from the definition of 
sins. 

Example 3. Compute log (1 + t). 

Since 1 4- » ■ \/2 [cos (ir/4) 4- i sin (t/4)], 


log (1 + *) - Log V2 + ( J + 2kr ^ i, fc - 0, ±1, ±2 


by (3-10). The principal value is got by setting k • 0. 
Example 4. Compute 2*. 

By (3-12), 

2 * - e iU *\ 

1 See Probe. 7 and 8. 



m 


SBC. 3} ANALYTIC ASPECTS 

But log 2 m Log 2 4- i2itk. Hence 

2 i - «**•*-** k - 0, rfcl, 4=2, . . .. 
Example 5. Compute t*. 

By (3-12), 

t* - «<**•*. 

But log i «■ Log 1 4- *[(*72) + 2/cr] « *K*V2) 4* 2 Jct], and hence 
{i „ c -<*/2-H*iO * - 0, dfcl, sfc2, . . .. 

Example 6. Find all solutions of the equation cos * — 2 *■ 0. 
We have cos a ** 2, which gives, successively, 



e “ + e_ “„2 

2 2 ' 


«“ + «-** - 4, 


«“* - 4«“ + 1 - 0. 

Solving for e u . 

. 4rfcVl6-4 


2 


a 2i x/3. 

Hence 

iz "* log (2 ± \/3 ) 

and 

a « \ log (2 rfc \/3 ). 


t 


Since log (2 db \/5 ) is infinitely-many-valued, there are infinitely many values of z. 

PROBLEMS 

1. Verify the following: (a) e ** ( b ) sin 2 a 4- cos 2 a — 1; (c) cos (aj 4- aj) 

* cos z\ cos tt — sin Zi sin z^; ( d ) cos iz « cosh a; (e) sin iz » t sinh a. 

2. If a and b are real integers, show that (re ie ) a +6t «* ^“^[cos (ad 4* 6 Log r) 4- 
t sin (ad 4* b Log r)J. 

3. Compute (a) oos (2 4* t), (b) 1*, (c) (1 4- 0\ (d) 2 1+ \ ( e ) 

4. Express in the form a 4- bi t where a and 6 are real: (a) l/(a — 1), (6) 1 /(«* 4. t), 
(c) sin (1 4- 1), (d) e 1 *, (e) e l/ *. 

6. Find the principal values and represent the numbers graphically: (a) log (—4), 

(b) log (5i), (c) log (1 4- *), (d) log t, (e) i\ (/) 6 1+ \ (g) sin 2i. 

6. Find all solutions of the following equations: (a) e* 4- 1 =* 0; (6) sin z — 2 «* 0 ; 

(c) cos'" 1 a * 2; (d) cos z— 1 *■ 0. 

7. The inverse functions are defined as solutions of the equation z * f(w) for w in terms 
of 2. Thus, w ** sin” 1 z if a — sin 10 » (e tw ~ e~ iw )/2i. Obtain e iv> in this example 

by solving the equation e 4< " — 2ize iw — 1 « 0. The Result is e tw ** iz db Vl — a*. 
Hence w •» sin"" 1 a «■* — i log (iz 4= VT ~ z* ). Show in the same way that 

tan 1 2 «• - log 

2 % — a 


and cos"* 1 % » — » log (« 4= \A* — 1 ). 



540 


[chap. 7 


COMPLEX VARIABLE 

*• Bofer to Prob. 7 and show that: 

(«) sinh -1 * «• log (* + V** + 1 ), (6) cosh -1 * — log (* + V** — 1 ), 

(c) tonh -1 1 m i log . 

2 1 **■ t 

9* For complex numbers a, b, c in what sense and in what circumstances is it true 
that (ab) e - a e b e ? 

4. Analytic Functions of a Complex Variable. We say that a point 
z « x + iy approaches a fixed point z 0 * x 0 + iyo if x x 0 and y y 0 . 
Let f(z) be a single-valued function defined in some neighborhood of the 
point z «= Zq. By the neighborhood of Zq we mean the set of all points in a 
sufficiently small circular region with center at z 0 . As z — > z 0 , the function 
f(z) may tend to a definite value w 0 . We say, then, that the limit of f(z) 
as z approaches z 0 is w 0 and write 

lim f(z) m w 0 . 

* > *0 

In particular, if f(z 0 ) = w 0 , we say that f(z) is continuous at z * zq. 

It is not difficult to prove that if f(z) — u(x,y) + iv(x,y) is continuous 
at Zo « % + iyo, then its real and imaginary parts u and v are continuous 
functions at (x 0 ,^o), and conversely. 

Let w = f(z) be continuous at every point of some region in the z plane. 
The complex quantities w and z can be represented on separate complex 
planes, called the w and 2 planes. The relationship w = f(z) sets up a 
correspondence between the points (x f y) in the z plane and the points 
(u } v) in the w plane (see Figs. 6 and 7), so that the corresponding points 
(u f v) fill some region R' in the w plane. 



if zo « Xo + iy 0 and z * zo + Az are two points in the z plane with 
* r ** + * Ay ’ the corres P° nd ^g Points in the w plane are w 0 » «o + iv 0 
and v> - wo + Aw, where Aw m Aw + t Av. The change Aw in the 






$41 


SEC. 4} ANALYTIC ASPECTS 

value of Wo <•* /(to) corresponding to the increment Az in to is 
Ate * f(to + A z) - /(to) 

and we define the derivative dw/dz [or/'(z)] by a familiar formula 


/'(so) 


Aw 

lim — 

At — • 0 A z 


lim 

At — ♦ 


/(z 0 + Az) - /(z 0 ) 
o Az 


(^ 1 ) 


It is most important to note that in this formula z « Zo + Az can assume 
any position in the neighborhood of z 0 and Az can approach aero along 
any one of the infinitely many paths joining z with z 0 - Hence, if the 
derivative /'(z<>) is to have a unique value, we must demand that the limit 
in (4-1) be independent of the way in which Az is made to approach zero. 
This restriction greatly narrows down the class of complex functions that 
possess derivatives. 

For example, if 

w Zly 

then on replacing z by z + Az and l by l + Az, we get 

w + Aw » (z + Az) (z -f Az) «= zl + z Az + z Az + Az Az, 


Hence Aw — zAz + zAz+AzAz 


Aw Az — 

and — - 2 + z b Az, (4-2) 

Az Az 

We show next that this quotient, in general, has no unique limit as Az 
is made to approach zero along different paths. Since z «* x + iy, 

Az = Ax 4* i Ay, Az » Ax — i Ay 

and we can write (4-2) as 

Aw Ax — i Ay 

— « x — iy -f (x + iy) b Ax — i Ay . (4-3) 

Az Ax + i Ay 

If we now let Az in (4-3) approach zero along the path QRP (Fig. 8), so 
that first QR ** Ay — » 0 and then PR « Ax 0, we get 

Aw 

lim — = 2x. 

&»-+ o Az 

But if we take the path QR'P and first allow QR' = Ax — » 0 and then 
R'P » Ay — > 0, we obtain 

Aw 

lim — * —2 iy. 

At -+ o Az 



542 


COMPLEX VARIABLE 


[CHAP. 7 


Except for x * y ■* 0, these limits are distinct, and hence w «• zl has no 
derivative except possibly at z ® 0. As a matter of fact, it is possible to 
show that this function does have a derivative (whose value is zero) only 
at the point z * 0. 



On the other hand, if we consider 

w — z? 


then w + Aw? « (z + A z) 2 * z 2 + 2z Az + (Az) 3 , so that 

Aw 2 z Az + (Az) 2 

— =. — 2 z + Az. 

Az Az 


The limit of this quotient as Az 0 is invariably 2z, whatever may be 
the path along which Az — > 0. In this example the derivative exists and 
its value is 2z. 

We obtain next a set of conditions which real and imaginary parts of 


w = f(z) as u(x,y) + iv(x,y) 

must fulfill if f(z) is to have a unique derivative at a given point z ~ x + iy. 
Since Aw *= Au + i Av and Az = Ax + i Ay, we get from (4-1) 


/'(*) 


Au + i Av 

= lim 

Az 0 Az 


Au + i Av 

= lim 

ax o Ax + i Ay 

Ay 0 


(4-4) 


Now, if we let Az 0 by first allowing Ay -* 0 and then Ax —* 0, 
we get from (4-4) 



(«) 



t 


SEC. 4] ANALYTIC ASPECTS 

If, on the other hand, we compute the limit in (4*4) by making first Ax 
and then Ay —► 0, we obtain 



543 

-4 0 


(4-6) 


Hence, if the derivatives in (4-5) and (4-6) are to have identical values at 
a given point z for these two particular modes of approach of A z to aero, 
we must have 

du dv dv du 

— - — * — (4-7) 

dx dy dx dy 


Equations (4-7) are known as the Cauchy-Riemann equations , and the 
foregoing calculation shows that they constitute necessary conditions for 
the existence of a unique derivative of f(z) = u(x y y) -f iv(x,y) at z =* x + ty. 
These equations also turn out to be sufficient 1 if one f urther assumes the 
continuity of partial derivatives in (4-7) at the point (x,y). 

Complex functions which have derivatives only at isolated points in the 
z plane are of minor interest in applications in comparison with those that 
have derivatives throughout the neighborhood of the given point. We say 
that a function f(z) is analytic (or holomorphic) at a given point z » z 0 if it 
has a derivative f'(z) at z = Zq and at every point in the neighborhood of z 0 . 
It can be shown that the following theorem 2 is true. 

Theorem. A necessary and sufficient condition forf(z) = u(x,y) + iv(x $ y) 
to be analytic at z 0 = Xo + Wo is that u(x } y) and v(x,y) together with their 
partial derivatives be continuous and satisfy Eqs. (4-7) in the neighborhood 
of (x 0 ,y 0 ). 

The points of the region where f{z) ceases to be analytic are called 
singular points of f{z). 

It is easy to show that familiar rules for differentiating sums, products, 
and quotients of real functions remain valid for analytic functions.* Also 
the formulas for differentiating elementary complex functions, defined in 
Sec. 3, are identical with the corresponding formulas in the calculus of 
real variables. We give a derivation of several such formulas in the follow- 
ing examples. 4 

1 A demonstration of this is given in several standard texts. See, for example, E. C. 
Titchmarsh, "The Theory of Functions/’ 2d ed., p. 68, Oxford University Press, London, 
1939. 

* This theorem can be deduced with the aid of the strong form of Cauchy’s theorem 
stated in Sec. 5. 

* See Prob. U 

4 See also Prob. 2. 



644 


COMPLEX VARIABLE 


(chap, 7 


Example 1. Show that de*/dz m e*. 

H xjo m e* a* 4*+% then the definition (3-1) yields 

w « u + iv » e*(cos y -f * sin y). 

Her®, t* ■■ e* cos y, t) * e* sin y, and it follows that 


du ** 

— « er cos y, 

dx 


du 

dy 


— s* sin y, 


— = rsmy, 


* , 

— “ c cos y. 
dy 


Since Eqs. (4-7) are satisfied and the partial derivatives are continuous, dw/dz can be 
calculated with the aid of either (4-5) or (4-6). Then, 

dw 

— « e* cos y -f ic* sin y 


*> e*(cos y + i sin y) ** e*. 

Example 2. Show that (d log z)/dz *» 1/z if z ^ 0. 

The function w » log z, as noted in Sec. 3, is multiple- valued. However any branch 
of this function got by fixing the value of k in (3-10) is single- valued, and the application 
of Cauchy-Riemann equations (4-7) to it shows that it is an analytic function except at 
z m 0. On fixing k we get from w * log z a single- valued function 

z - e w 


whose derivative with respect to w by Example 1 is 


Hence 



dw d log z 1 
dz dz z 


if t & 0. 


The point z *■ 0 is a singular point of w « log z, since the derivative at that point 
ceases to exist. 

Example 3. Show that dz n /dz » nz n ~ l for all values of n (real or complex). 

If w *• z w , then 

log w m n log z. 

On differentiating this with respect to z, we get 

1 dw n 

w dz z 

.< dw w 

Hence ~ • n - - nz*~\ 

at % 

since w m z n . This derivative oeases to exist at * «* 0 if n <1. 



SBC. 5] 


ANALYTIC ASPECTS 


545 


PROBLEMS 


1* Show that 

(O) ~(fx =fc/ s ) -//(*) ±/&), W |(/iA) - A/2 +/j/J. 


to 


d* 


d) 




whenever /i and h are analytic functions. 
2. Show that 


(a) 

(d) 


d(cos z) 


dz 

<J(iaiT 


«* — sin z t (b) 
'z) 1 


d(sin z) 
dz 


cos z, (c) 


dfidh 
df% dz 1 

d(tan z) 
dz 


dz 


1 Z‘ 


d(sinh z) /rs da* 

- > ( € ) -- » cosh z , (/) 


dz 


dz 


** sec 2 *, 

** a* log a. 


3. Determine where each of the following functions fails to be analytic: (a) z 2 -f- 2 z 
(b) z/(z -f 1), M 1 /* + (z ~ l) 2 , (rf) tan z , (e) l/[(z - l)(z -f 1)],(/) 2 | > (^) £ ) *2 - J 

- (i) x/(x 2 -f y 2 ) -f %/(* 2 + I/ 5 ), 0) |*|, W tan" 1 z , 

6. Integration of Complex Functions. Cauchy’s Integral Theorem. We 

define the integral J c /( z ) dz of a complex function f{z) = u(x, 2 /) + 
along a path C in terms of real line integrals as follows: 


f f(z) dz s (u + zV)(dr + t dy) 


= J c (u dx - V dy) + if c ( vdx + u dy). ( 5 - 1 ) 

Real integrals of this type were studied in Chap. 5, Sec. 4, where it was 
observed that they exist when the functions u(x y y) and v(x,y) are continu- 
ous and the path C is sufficiently smooth. 

The integral in (5-1) can also be defined in a manner of Sec. 4, Chap. 5, 
by the formula 

r n 

/_/(*) dz - lim 23 /(f.)(z. - z«-i). (5-2) 

J{u n 

xnxx\t i ~t i _ l \ -> 0 

It is supposed that the curve C is divided into n segments by points z>- 
and that f* is some point of the ith segment. The limit is then computed 
as the number of segments is allowed to increase indefinitely in such a 
way that the length of the largest segment tends to zero. The fact that 
the definitions (5-1) and (5-2) are equivalent follows from consideration 
of Sec. 4, Chap. 5. 

As an illustration of the use of formula (5-1) consider the integral 



54$ COMPLEX VARIABLE [CHAP. 7 

where the path C is a straight line joining the points 2 = 0 and z = 1 *f 
2 1 (Fig. 9). Since l 2 = (x - iy) 2 = x 2 - y 2 - i2zy, we get, on sub- 
stituting u ~ x 2 — y 2 ,v =* —2xy in (5-1), 

jf s 2 dz = ^ [(r 2 - y 2 ) dx + 2xy dy] + if c [ — 2xy dx + (a: 2 — y 2 ) dy]. 

(5-4) 



But the cartesian equation of C is y = 2x, and hence (5-4) can be reduced 
to the evaluation of two definite integrals: 

jz 2 dz = 5x 2 dx + — 10x 2 dx = % — i x %. 

The value of the integral (5-4) depends on the path C joining the given 
points z = 0, 2 = 1 4- 2i, for according to Sec. 9, Chap. 5, a necessary and 
sufficient condition that the line integral 


f Mdx + Ndy 
Jc 


(5-5) 


be independent of the path in a simply connected region R is that 


dM _ dN 
dy dz 


(5-6) 


throughout R. We further recall that in deducing the condition (5-6) 
with the aid of Green’s theorem it was supposed that M(x,y), N(x f y), 
and their partial derivatives in (5-6) are continuous functions throughout 
the region. It is readily checked that Eq. (5-6) is not satisfied by the 
functions appearing in the line integrals in (5-4). 

If, however, f(z) = u + iv in (5-1) is an analytic function, then the 
Oauehy-Riemann equations (4-7) demand that 


du dv dv du 

dx dy dx dy 


(5-7) * 



ANALYTIC ASPECTS 


547 


SEC. 5] 

Reference to (5-6) shows that these conditions are precisely those that 
ensure the independence of the path of the line integrals in (5-1), provided 
that the partial derivatives in (5-7) are continuous functions in the given 
simply connected region R. Thus, if we suppose that f(z) is analytic in the 
given simply connected region and /'(z) is continuous there, then the 
integral 

f c f{z)dz 

is independent of the path joining any pair of points in the region. If 
the path C is closed, then the value of this integral is zero. We thus 
have a theorem, first deduced by Cauchy, which is of cardinal importance 
in the study of analytic functions. Although the foregoing proof assumes 
the continuity of /'(z), the theorem can actually be established 1 under 
the sole hypothesis that f(z) exists at each point of the region, and we 
state it in this strong form. 

Cauchy’s Integral Theorem. If f (z) is analytic at all points within 
and on a closed curve C, then dz ~ 0. 

We conclude this section by deducing, from definition (5-2), a useful 
inequality furnishing an upper bound for the value of the complex integral 

j c f( z ) dz. Inasmuch as the modulus of the sum of complex numbers is 
never greater than the sum of the moduli, 

I { c f(z)dz\< f c \f(z)\-\dz\. 

Now, if the modulus |/(z)| of f(z) along C does not exceed in value some 
positive number M f then 

j dz j < M (dz| = M | dx + idy | = M j^ds - ML (5-8) 

where L is the length of C . 

As an illustration of the use of the inequality (5-8) we apply it to deduce 
an upper bound for the integral (5-3) . The modulus of l 2 takes its maximum 
at the point z = 1 + 2 i. Hence we can take M in (5-3) as ] I -f 2i\ 2 = 5, 
and (5-8) then yields 

| j z 2 dz J < 5^/5, 

inasmuch as L ** V5 for the rectilinear path in (5-3). 

1 See Titchmarsh, op. cit. $ pp. 75-83. In a somewhat different development of the 
subject one deduces the continuity of f\z) from Cauchy’s theorem, not the other way 
about. We shall see in Sec. 7 that the theorem actually implies existence and continuity 
of derivatives of all orders. 



548 


COMPLEX VARIABLE 


{CHAP. 7 


PROBLEMS 


1. Find the value of the integral f z 2 dz along the rectilinear path joining the points 

Jc 

* m 0 and * • 2 + t. Show that this integral is independent of the path. 

1 Find the value of the integral l l dz along the rectilinear path y » x joining the 

Jc 

points (0,0) and (1,1) and also along the parabola y « x 2 joining the same points. 

3. Show that the integral I z dz evaluated over the path | z | ** 1 in a counterclock- 

Jc 

wise direction yields 2iri. Note that z — e* and l ** along the path |*| « 1. 

r l 2 — i 

4. Find the value of the integral / dz , where the path is the upper half of the 

J- i ^ 

circle |*| *■» 1. Calculate the value of this integral over the lower half of the circle 

5. Show that / (1 -f * 2 ) ds is independent of the path C, and evaluate this integral 

Jc 

when C is the boundary of the square with vertices at the points « ■ 0,« » l,i ■ 1 + *, 
and z * i. 

6. What is the value of the integral I e % * dz where C is the boundary of the square in 

Prob. 5? ° 

7. Find the value of the integral / e* dz over any path joining * «■ 0 and * - vi. 

J o 

8. Use formula (5-8) to show that: 

1 j { r2+t rj, | | rt i 

(a) J j z l dz J < 10, (6) j J “j j < 2, (c) J J (x 2 + *t/ 2 ) dz J < 2, 

where paths are straight lines joining the points appearing in the limits of these integrals. 


6. Cauchy’s Integral Theorem for Multiply Connected Regions. In 

establishing Cauchy’s integral theorem in the preceding section we assumed 





Fig* 10 


that the region bounded by the 
curve C is simply connected. It is 
easy to extend this theorem to 
multiply connected domains in the 
maimer of Sec. 9, Chap. 5. Thus 
consider, for definiteness, a doubly 
j connected region (Fig. 10) bounded 
/ by closed curves C\ and C 2 , where 
' C 2 lies entirely within C x . We 
assume that f(z) is analytic in the 
region exterior to C 2 and interior to 
Ci and analytic on C 2 and C x . The 
requirement of analyticity on C\ and 
C 2 implies that the function f(z) is 


analytic in an extended region (indi- 
cated by the dashed curves K x and K 2 ) that contains the curves Cj and C 2 . 
If some point A erf the curve C x is joined to a point B of C 2 by a crosscut 



SEC. 6] ANALYTIC ASPECTS 549 

AB, then the region becomes simply connected and the theorem of Cauchy 
is applicable. Integrating in the positive direction gives 

<f APA /(*) dz + j Ag m dt + <fi BQB /(«) dz + j BA f(z) dz - 0 , (&- 1 ) 

where the subscripts on the integrals indicate the directions of integration 
along Ci, the crosscut AB, and CV Since the second and the fourth 
integrals in (6-1) are calculated over the same path in opposite directions, 
their sum is zero and one has 

<£ c f(z) dz + (j> c f(z) dz ® 0 , ( 6 - 2 ) 

where the integral along C\ is traversed in the counterclockwise direction 
and that along C 2 in the clockwise direction. Changing the order of 
integration in the second integral in (6-2) gives 

<j> c f(z) dz » fix) dz. (6-3) 

We see that the values of the integral of f(z) over two different paths 
Ci and C 2 are equal, but they need not be zero inasmuch as J{z) may not 
be analytic at every point of the region bounded by GV But whatever 
may be the value of the integral over the path C 2 , it is the same as its 
value over the path CY An important principle of the deformation of 
contours follows at once from this observation: The integral of an analytic 
function over any closed curve C\ has the same value over any other curve C 2 
into which C\ can he continuously deformed without passing over singular 
points of f(z). 

We shall see that this principle will enable us to simplify the computation 
of integrals of analytic functions. 

The foregoing results can be extended in an obvious way to yield the 
following theorem: 

Theorem. If f (z) is analytic in a closed multiply connected region hounded 
by the exterior curve G and the interior curves C u C 2} . . G«, then the in- 
tegral over the exterior curve G is equal to the sum of the integrals over the 
interior curves provided that the integration over all the contours is performed 
in the same direction. 

It should be noted that the requirement of analyticity of f(z) in the 
closed region implied that }{z) be analytic on all contours forming its 
boundary. 

Before considering applications of the theorem of this section to specific 
problems, we deduce an important result which will enable us to compute 
many integrals by a method which is vastly simpler than that developed 
in Sec* 5* 



COMPLEX VARIABLE 


m 


[chap. 7 


7. The Fundamental Theorem of Integral Calculus. Let f(z) be analytic 
in a simply connected region R (Fig. 11), and let C be a curve joining two 

points P 0 and P of the region determined 
by the complex numbers zq and z. We con- 
sider the integral 



/7(*)* 

Jea 


(7-1) 


along C. Since /(z) is analytic, the integral 
(7-1) is independent of the path, and its 
value is completely determined by the choice 
of z Q and z. If z 0 is fixed, the integral (7-1) 
defines a function 

F(z) = f‘m dz (7-2) 

J «0 

for every choice of z in R. 

To emphasize the fact that the integration variable z plays a distinct 
role from the variable z appearing in the upper limit of the integral, we 
can rewrite (7-2) as 


m 


rm«. 

J t 0 


(7-3) 


We prove next that F(z) is an analytic function and, moreover, its 
derivative at any point z has the value of the function in the integrand 
at that point. That is, 

F'(z) «/(*). 

We can use (7-3) to compute the difference quotient 
F{z + A z) — F(z) 

Az 


A z 


i[C‘'K)«-ij<s)«] 

1 r*+Aa 

- /, m «, (M) 


and rewrite (7-4) by adding and subtracting /(z) in the integrand: 
F(z -f- Az) — F{z) 1 r*+A» 

- — — — — — - ■ ■ sat 

Az Az 

1 

* Az" ^ A z 


[/(f) -f(z) + /(z)]df 


/•*4-Aa 1 ft 4-A* 

;[/(*)[ <*fl + -/ (/(f) 

! J * Az Jz 


m* t. 



ANALYTIC ASPECTS 


651 


SBC. 71 

d$ » Az, so that 


Now if 


F(z + Az) ~ F(z) 
Az 


A*) + ~ f +A ‘ t/(f) -/(»)!*. 

Az •'* 


lim ~ r +A '[/(f) -/(*)] dr -0, 

A»-+0 Az 


( 7 - 5 ) 

( 7 - 6 ) 


then it would follow from (7-5) that F'(z) « f(z). The fact that the limit 
in (7-6) is, indeed, zero follows at once from the estimate (5-8), for if 
M « max | f(t) — f(z) | on the path joining z and Az , then 


1 re- 

— / i/(f) -midi 

Az 


< M. 


But since f{z) is continuous, M — > 0 as Az ~ > 0. 

Any function Fj(z) such that F\(z) «= f(z) is called a 'primitive or an 
indefinite integral of f(z) . As in real calculus, it is easy to prove that if 
Fi(z) and F 2 (z) are any two indefinite integrals of /(z), then they can 
differ only by a constant. 1 

Hence, if F x (z) is an indefinite integral of /(z), it follows that 
F(z) « ff(z) dz = F x (z) + C. 

f* 0 

To evaluate C, set z * z 0 ; then, since / f(z) dz » 0, C = — F^zq). Thus 

-'do 

F(z) = /*/(*) de = F,(z) - F,(*o). (7-7) 

■/ZO 


The statement embodied in (7-7) establishes the connection between line 
and indefinite integrals and is called the fundamental theorem of integral 
calculus because of its importance in the evaluation of line integrals. It 
states that the value of the line integral of an analytic function is equal to 
the difference in the values of any primitive at the end points of the path of 
integration . 

1 Proof: Since F[{z) » F£(z) « /(*), it is evident that 


F[(z) - Fl(*) 


d(F x - F 2 ) dO n 

23 m 0 . 

dz dz 


But if dG/dz « 0, it means that G\z) « (du/dx) *f i(dv/dx) *» (dv/dy) — i(du/dy) ■» 0, 
so that du/dx ** dt>/dx « du/dy - *■ 0, and thus v and t> do not depend on x 

and y. 



582 COMPLEX VARIABLE fCEAP« 7 

Example I. As an illustration of the use of formula (7-7) consider the evaluation of 

/>d* (7-8) 

Jc 

along some path C joining z • 0 and * » 2 *f i. Inasmuch as /(a) * a* is analytic 
throughout the finite z plane, the integral (7-8) is independent of the path. Moreover, 
Since F(f) — is an indefinite integral for f(z) — z 2 , we can write 

I 2 * 1 1 


r 


z 2 dz * - z 3 I 


:(2-M) 3 . 


The reader should contrast this computation with calculations required for solving this 
in Prob. 1, Sec. 5, 

Example 2. Evaluate I e* dz over some path C joining z - 0 and t m vi. Since e* is 

Jc 

analytic, we get at once from (7-7) 


r 


e* dz 


- 1 - - 2 . 


We indicate the nature of required calculations if this integral were to be computed 
by the method of Sec. 5. We first separate the integrand into real and imaginary parts, 

s* m e • e* cos y + ie* sin y, 

and form two real line integrals 

I e* dz «* / (e* cos y + ie* sin j/)(dx -f i dy ) 

Jc Jc 

«■» j (e*coBydx — z* sin y dt/) -f * f (e* sin y dx «f e* cos y dy). 


Since these line integrals are independent of the path, they may be evaluated over any 
convenient path joining the points (0,0) and (0,*-) corresponding to z «* 0 and z « x i. 
The result of such calculations would yield -2, as the reader can verify. 

Example 3. Discuss the integral J (z — a) m dz, where m is an integer and a is a 
constant. 

The function f(z) ~ (z ~ a) m is obviously analytic at all points of the z plane as long 
as m is a positive integer. If m < 0, we write m * -n and consider 


/(*) 


1 


where n is a positive integer. 


(z ~ a) n 


To evaluate 


f (z — a) w dz for m > 0, we note that 

*° (z — a) w+1 

P(i) - V ' 


m -f 1 

is an indefinite integral for /(z) =* (z - a) m , Accordingly 

A. m + I L 


(7-9) 


(7-10) 


If, in particular, the path C is dosed, so that the limits in (7-10) coincide, we conclude 
that the value of the integral is zero. This result also follows from Cauchy’s theorem, 
since /(*) « (* — a) m isranalytie for all values of z when m > 0. 



SEC* 7 ] ANALYTIC ASPECTS 

We consider next the integral 


683 


r dz 
Jc (« — * 


and note first that if the path C passes through the point z » a, the integrand becomes 
meaningless at z ** a. In this book 1 we shall not consider in detail integrals over those 
paths that go through singular points of the integrands, but special types of such inte- 
grals will occur in Sec. 22. 

If C is a closed path and a is not in the region R enclosed by C, the integrand in (7-11) 
is analytic in the closed region R, Hence, by Cauchy's theorem the value is zero. If, 
however, a lies in R , Cauchy’s theorem 
does not apply, since f(z) ** l /(* — a) n u? 
ceases being analytic at z ■» a. The inte- f 

gral (7-11) can, of course, be evaluat'd by I \ 

the method of Sec. 5 once the equation of y \ 

C is specified. However, it is wise to sim- "N \C 

plify calculations by making use of the ) s — \ 

principle of deformation of contours. This J ( s\ \ 

principle states that when z — a is in the / V a y \ 

interior of C, f N | 


r c (z~a) n T y (z-a) n 


where 7 is a circle with center at a and 

with radius p so small that 7 lies within C "q 

(Fig. 12) But the integral over 7 is easily 
evaluated. Setting z — a ** pe l6 , we get 
dz = pe^i dd on observing that p is constant on 7. Hence 



y dz y pe' 6 i d$ _ i f' 1 
Tc ” Ty * p"- 1 Jo 


r c (z-~ar t 

If n ** 1, we get 


e (l-n)0i ^ 


g(l-~n;(h 2*- 

1 t(l - n) 0 


if n it 1. 


I — -if 

Jc z — a Jo 


dO « 2 iri. 


In evaluating the integral (7-12), we noted that the integrand e (i ~ n)et dd \ for n 5^ 1, is 
the differential of e a ~ n)9% /i(\ ~ n), and we made use of the fundamental theorem of 
integral calculus. 


J r dz 

~r , where C is the circle x 2 4* y 2 * 4. 

c * — 1 


The function 


m 


1 _ 1 

~ 1 % - 1)(* + 1) 


(7-14) 


1 When z » a lies on the path of integration, the integral in (7-11) is an improper 
complex integral and it calls for special considerations analogous to those required to 
treat improper real integrals. Certain types of improper complex integrals are of in- 
terest in applications. See, for example, N. I. Muskhelishvili, “Singular Integral Equa- 
tions/’ P. Noordhoff, N,V., Groningen, Netherlands, 1953. 



COMPLEX VARIABLE 


554 


fCHAP. 7 


has two singular points t — 1 and «■»—!, both of which He within the given circle 
|*| < 2 (Fig. 13). If we delete these points from the circular region <7 by circles 71 and 
of sufficiently small radii, /(*) will be analytic in the triply connected domain exterior 
to 7 i and 7 * and interior to C. Then Cauchy's theorem for multiply connected domains 
permits us to write 

f /(*) dz~[ /(*) d* + f m (MW 

Jc jy ! Jyt 



The integrals in the right-hand member in (7-15) are readily evaluated. Since 


we get 


1 11 11 
(z- 1 )(z + 1) “ 2 z - 1 ” 2 z~+T 


L 


1 


(Z - 1 )(* + 1) 



1 f dz 

2 J yi z + 1 


(7-16) 


If the radius of 71 is such that 71 contains within it z * -f 1 but not z « — 1 , then 


by (7-13), and 


L 

L 



dz 

z + 1 


2m f 

0, 


by Cauchy’s integral theorem, for l/(z + 1 ) has no singularities within 71 . 
first integral on the right in (7-15) has the value n. An entirely similar 
shows 


Therefore, 


ly , (* ~ 1 )(* + 1 ) ^ 

Ic {* - IK* + l ) d * " M 


Thus, the 
calculation 


even though the integrand is not analytic in the region | z j < 2. 



SEC. 8 J 


ANALYTIC ASPECTS 


f 


555 


PROBLEMS 

1. Show that f zdz ** %(& — sg) for all paths joining zq with *. 

-'so 

2. Evaluate the integral I (z a)*” 1 dz, where C is a simple closed curve and a 

Jc 

is interior to C, by expressing it as a sum of two real line integrals over C. Hint; Set 
z — a «■ pe**; then dz «* e? l (dp 4" ip d$). 

3. Evaluate / z~ 2 dz where the path C is the upper half of the unit circle whose 

Jc 

center is at the origin. What is the value of this integral if the path is the lower half of 
the circle? 

4. Evaluate / z~~* dz, where C is the path of Prob. 3. 

Jc 

6. Evaluate / ( z 2 — 2z -f 1) dz, where C is the circle x* -f V 1 ** 2. 

Jc 

r z 4 1 

6. Discuss the integral / — =— dz , where C is a path enclosing the origin. 

Jc 2 

7. What is the value of the integral / (1 -f z 2 )^ 1 dz, where C is the circle x* -f y* m 

Jc 

8. Discuss Prob. 7 by noting that 

1 


9 ? 


1 ■+- 2 2 


I(i U 

2 i\z — t z 4 * i/ 


and evaluating the integrals over the unit circles whose centers are at * — i and z «■ — i. 
Note the theorem of Sec. 6. 

9. Show that the integrals (a) f • - , (b) f sin z dz, (c) f ze* dz, (d) f z~ 2 dz vanish 

Jc z — 2 Jc Jc Jc 

if C is the unit circle \z \ « 1. 

10. Evaluate the integral f — - — dz along the following paths C: (a) \z\ ** 

Jc 1—2 

(b) \z | “ 2, (c) \z — 1 1 « 1, (d) \z 4* 1 1 « 1. Hint: Decompose the integrand into partial 
fractions as in Prob. 8. 


8. Cauchy’s Integral Formula. In this section we deduce with the aid 
of Cauchy's theorem the remarkable fact that every analytic function /(s) 
is completely determined in the interior of the given closed region R when 
the values of f(z) are specified on its boundary. 

Let f{z) be analytic in a simply connected region R and on its boundary 
C. If a is an interior point of R , then the function 


m 

z — a 


(8-1) 


is analytic in R with the possible exception of the point z * a. If this 
point is excluded from the region by enclosing it in a circle y of radius p 
and with center at a (Fig. 12), then (8-1) will surely be analytic in the 
region exterior to y and interior to C. 



COMPLEX VARIABLE 


[chap. 7 


It follows, then, from (6-3) that 

m , f m 

' z 


J c z ~ a J y i 


dz 


(8-2) 


where the paths (7 and y are described in the same sense. Now the integral 
in the right-hand member of (8-2) can be written as 


*y z — a J y 


m -m 


z — a 


dz + f (a) — 


dz 


But by (7-13) 


ly z 


dz 


= 27 ri y 


(8-3) 


(8-4) 


and we shall show next that the first integral on the right in (8-3) has the 
value zero. Indeed, if we take z — a = pe %e , then, as long as z is on 7, 
dz = ipe 10 d$ y and therefore 


/ / (2 ) /( - a) dz = if \}(z) - /(a)] dd. 

J y z ~ a J y 


Let the maximum of | f(z) — f(a) | be M; then by (5-8) 


L 


f(z) - f(a) 


dz 


/•2t 

< M / d$ 

Jo 


2rM. 


(8-5) 


(8-6) 


The radius p is arbitrary, and if we make it sufficiently small, then 
max |/(z) — /(a) | can be made as small as we wish, since f(z) is a continuous 
function. Accordingly, M — ► 0 as p — ► 0 On the other hand, from the 
principle of deformation of contours, the value of the integral (8-0) is 
independent of the radius p. Since M —> 0 when p — * 0, we conclude 
that the value of the integral (8-5) is zero. 

Accordingly, (8-3), together with (8-4), gives the result 


f /(«) 

Jc z — a 


dz = 2t if (a). 


(8-7) 


We recall that the point a is any interior point of the region R bounded 
by C and z is the variable of integration on the contour C. If we denote 
the variable of integration by f and let z be any interior point, we can 
rewrite formula (8-7) as 


/(*) 


( md * 
2ri f — z 


(8-8) 


Formula (8-8) permits us to calculate the value of f(z) at any interior 
pbint from specified boundary values /({*) on the contour C. It is known 



ANALYTIC ASPECTS 


SEC. 8} 


557 


as Cauchy’s integral formula. This formula can be extended in the man mr 
of Sec. 6 to multiply connected domains bounded by the exterior contour 
Co and m interior contours C i, C 2 > . . . > C m . The integration in (8-8) is 
then performed in the clockwise sense over the interior contours and 
counterclockwise over the exterior contour Cq. 

It is not difficult to show with the aid of formula (8-8) that an analytic 
function f(z) has not only continuous first derivatives in the region but 
also derivatives of all orders. Thus an analytic function can be differen- 
tiated infinitely many times. 

In fact, if we consider an integral of Cauchy’s type y 


F(z) 


1 

2 iri 


i h r 


/(f) 




(8-9) 


where /(f) is any continuous (not necessarily analytic) complex function, 
then this integral defines an analytic function F(z). To show this we 
merely have to prove that F(z) has a derivative at every point of the 
region R bounded by C. We form the difference quotient with the aid of 
(8-9) and get 


F f (z) 


lim 

0 


F(z + A z) — F{z) 
Az 


Dm ±[*>* 1 

az — * o Az L 2 iri f — (z + Az) 2 wi f — z J 

lim r Li — m* — 1. 

Az -+ 0 L27Ti 1C (f — Z — Az)({ — z) J 


On taking the limit as Az — ► 0 under the integral sign, which is legitimate 
if /(f) is continuous, we get 


r M 


-/ 

2 ri J (- (f 

Continuing in the same way, we find 


m 


z)' 


A- 


F"(z) 


21 
2 iri 


f /(f) 

JC (f - 2 )3 


df, 


F <n) (z) 


_»!_ f /(f) 

2tTI dc (f - 2)" +1 


df. 


We have thus shown that F(z) defined by (8-9) has derivatives of all 
orders even when nothing is said about the relation of the values of F(z) 
on the boundary C to the function /(f) appearing in the integrand. In 



COMPLEX VARIABLE 


558 


(chap. 7 


the special case when /(f) * F(f), we have a formula for the nth derivative 
of the analytic function /(z) at any interior point of R in terms of the 
values of /(z) on C : 


/<»>(*) 


nl r / 

2ri Jc (t - 


/(f) 


2 « J c (f - z) n+1 


d f, 


0 , 1 , 2 , .... 


( 8 - 10 ) 


We conclude this section by noting some important consequences of formula (8-7). 
Let. the path C be the circle 1 1 — a | «* p with center at z — a and with radius p. Sup- 
pose that the maximum value of the modulus of f(z) on this circle is M; then by (6-8) 


l/(a)l<~-2rp 
2 IT p 


M. 


This result is independent of the radius p. Consequently |/(z) | at the center a of the 
circle is not greater than its maximum value on the boundary. Using this result one can 
prove that if f(z) is analytic in a given region R bounded by a curve C, and if M is the 
maximum value of | f(z) | on C, then \f(z) | < M at each interior point of R unless 
|/(z) | «* M throughout the region. This result is known as the maximum modulus 
theorem .* The fact that |/(*)|< M follows from Sec. 24, Chap. 6, if we note that 
log \}{z) | is harmonic. 

Example 1 . Find the value of the integral f — - dz if C is the ellipse x 2 -f 4y 2 ** 1 . 

Jc z 

Since sin z is analytic in the region bounded by C, formula (8-7) yields, upon setting 
f(z) — sin z and a — 0, 


f sin z 
JC * 


Example 2. Evaluate the integral j 


Jcz + l 

The point z — — 1 lies within the given circle, and since e 
formula (8-7) yields 

f I 

/ — — dz — 2i tie ‘ — e2w i. 

Jc* 4- 1 U i 


dz over the circular path | z | » 2. 

is analytic within C, 


Example 3. Find the value of the integral 

tan z 


L 


TjCfe, 


Jc 1* ~ (ir/4)] 2 

where C is the circle |s| — 1. 

The point * - tr/4 lies within C , and we note that tan z is analytic for \z | < 1. From 
( 8 - 10 ) 

™-hh m 

Hence 


idz. 


2ri J (z — a) 2 

/ tan * . „ . fd tan z\ oT 

_____ * - 2« (— -j < - 2« sec® - - 4«. 


r «- T/4 


‘See proof, for example, in E. C. Titchmarsh, “The Theory of Functions,” 2d ed., 
p. 164, Oxford University Press, London, 1039. 



SEC. 9} 


ANALYTIC ASPECTS 


559 


t 


PROBLEMS 


r -f 1 

1. If f(z) «* J - <ff , where C is the circle of radius 2 about the origin, 

find the value of /(I — t). 

2. Apply Cauchy’s integral formula to Prob. 7, Sec. 7. Use the integrand in the 
form given in Prob. 8, Sec. 7. 

8. Evaluate the following integrals over the closed path C formed by the lines x • dfcl, 
t sin z F cos z F g* f 

y — ±1: (a) / dz, (6) / dz, (e) / — dz, (d) / (sin * + «•) d*. 

Jc 25 /C * JC 2 ~ 3^1 7c 

F cosh z , 

M / 

dr * 

4. Evaluate with the aid of Cauchy’s integral formula 


/• gr + r , f 

Jc? - 1 


where C is the circle |f | — 2. Hin/: Decompose the integrand into partial fractions. 

6. What is the value of the integral of Prob. 4 when evaluated over the circle | f — 1 1 
* 1? Hint: Note that (3f 2 + f)/(f + 1) is analytic for |£ — 1 1 < 1. 

/*3z 2 -f 2z “ 1 

8. Evaluate / dz, where C is the circle \z \ » 1. 

Jc 2 

7. Can j f(z) | as same a minimum value at an interior point of a region within which 
f(z) is analytic? Consider f(z) *• z. 

8. Can | f(z) | assume a nonzero minimum at an interior point of a region within which 
/(z) is analytic? Hint: Consider l//(z). 


9. Harmonic Functions. We saw in the preceding section that a function 
analytic at a given point of the region has derivatives of all orders at that 
point. It follows from this that the real and imaginary parts of an analytic 
function }{z) = u + iv have partial derivatives of all orders throughout 
the region where f(z) is analytic, for by (4-5) and (4-6) 


/'(*) 


du dv 

h i — 

dx dx 


and since f'(z) is also analytic, 


/"« 


d 2 u d 2 v 

l i 

dx 2 ^ dx 2 


d 2 v 


dv 

dy 

d 2 u 


du 

dy 


d z U 


d 2 V 


dx dy dx dy dy 2 dy 2 


The fact that } n (z) is analytic enables us to differentiate again to obtain 
the third partial derivatives, and so on. 

Inasmuch as the existence of the third partial derivatives ensures the 
equality of mixed partial derivatives of the second order, we can show 
that the real and imaginary parts of an analytic function satisfy Laplace's 
equation throughout the region of analytieity of f(z); for on differentiating 



560 COMPLEX VARIABLE [CHAP. 7 

tiie first of Cauchy-Riemann equations (4-7)with respect to y and the second 
with respect to x, we get 


b 2 u 

by bx 

and adding these we find 


b 2 v 

ay 2 ’ 


b 2 v 
bx 2 


b 2 V b 2 V 

b^ + by 2 


0. 


The fact that u also satisfies Laplace's equation 

0 


b 2 u b 2 u 
b^ + b^ 


d*U 

dx dy 


follows similarly from the differentiation of the first of Eqs. (4-7) with re- 
spect to x and the second with respect to y. 

Any real function u(x,y) with continuous second partial derivatives 
which satisfies Laplace’s equations in a given region is called harmonic 
in that region. Thus the real and imaginary parts of a function analytic 
in the region R are harmonic functions. Two harmonic functions u(x,y), 
v(x,y) such that u + iv is an analytic function f(z) are said to be conjugate 
harmonics. We shall show next that if one harmonic function is given, its 
conjugate harmonic can be determined to within a constant of integration. 
For, let u(x f y) be given in R. Then if v(x y y) is a conjugate harmonic, these 
functions satisfy the Cauchy-Riemann equations 



du dv 

du 

dv 




dx dy 

% “ 

dx 




dv 

bv 


du 

du 

Hence 

dv = — 

dx dy 

= — 

— dx H dv 


dx 

dy 


dy 

dx 

and, since du/dx and du/dy are know from u(x f y) f we have 



r(x,v) / 

du 


du \ 


v(x,y) 

= / ( - 

■ — dx 4* 

— dy jf 


* (*o*Vo) \ 

dy 


dx / 


where the integral can be evaluated over any path joining an arbitrary 
point (x 0 ,yo) of R with ( x,y ). Since the value of the line integral (9-2) 
depends on the choice of (xo>2/o)> *t 18 c ^ ear that v(x,y) is determined only 
to within an arbitrary constant. The integral is independent of the path 
inasmuch as 

by V by/ bx \bx/ 

and this equation is true because u(x,y) is harmonic. It should be noted 



ANALYTIC ASPECTS 


EC. W] 


mi 


hat when the region R is not simply connected, the function v(x,y) may 
um out to be multiple-valued. 1 

The connection of analytic functions with Laplace s equations is one 
& the principal reasons for the importance of the theory of functions of 
omplex variables in applied mathematics. 

In the preceding section we noted the maximum modulus theorem for 
inalytic functions. This theorem enables us to prove the important fact 
hat the maximum values of harmonic f unctions ( which are not mere constants ) 
ire invariably assumed on the boundary of the region . 

Let u be harmonic in the region R whose boundary is C. If v is a con- 
jugate harmonic, then u + iv is an analytic function, and therefore the 
function 

e u+%v _ e «( cog v -f- 1 sin v ) 


[s also analytic. But the maximum of \e u + tv \ ss e u is assumed on the 
boundary C of R by the maximum modulus theorem. Since e u takes on 
its maximum on the boundary 0, u(x,y) must assume its maximum on C. 


Example' The function u ** x 2 — y 2 is harmonic in every region. Obtain a conjugate 
harmonic v. 

Inserting u in the formula (9-2) yields 

Ax.v) A*,v) 

v(x t y) « / (2 y dx -f 2x dy) « 2 / d{xy) * 2 xy + c, 

•f(xo-Wo) ■'(aro. Vo) 

where c -~2xoyo. 

In this problem the integrand is so simple that we wrote its differential by inspection, 
hi a more complicated case it may prove more expedient to evaluate the integral over 
some convenient path rather than reduce the integrand to the form of a differential of 
some function. 


PROBLEMS 

1. Prove that v « 3 x 2 y — y 8 is harmonic, and find a conjugate harmonic u. 

2. Find an analytic function /(a:) * u 4- iv if: 

(a) u *» x; 

(b) u «* cosh y cosx; 

(c) u m x /(x 2 4* y 2 )i 

(d) u ** e* cos y ; 

(<?) u « log Vx 2 4- y 2 . 

10. Taylor’s Series. In this section we are concerned with the power- 
series representation of analytic functions. The reader is advised to review 

Secs. 8, 9, and 16 of Chap. 2 dealing with the properties of power series. 

00 

Here we recall that when the power series X a ^ k converges for z ® z lt 

ktO 

it converges absolutely and uniformly in every closed circular region 
\z\ < r, where r < |zi| . A circle of radius r such that Sa* 2 * converges for 

1 See Bee, 5, Chap. 5. 



COMPLEX VARIABLE 


562 


(chap. 7 


|*| < r and diverges for every \z\ > r is called the circle of convergence, 
and the number r is the radius of convergence . The radius of convergence 
can frequently be determined with the aid of the ratio test. Thus 


lim 


whenever this limit exists. 1 


a n -i 


00 ( — 1 ) n z n 

Example: The series — has the radius of convergence r 

n- 1 n 


1, since 


lim 


Qn-l 

On I 

The series 2 n k* converges only for z ■ 

n— 0 

Urn |2=i 


lim 


1 


1. 


n — * « n 

0, since in this case 
1 


lim 


► « n 


« 0. 


_ g 

On the other hand, the series 2J — - converges for all values of r, since 
»~o nl 


lim 

n —+ « 


1 

On 


«*> lim 


(n — 1) ! 

We saw in Sec. 9, Chap. 2, that with every real function /(x) having deriv- 
atives of all orders at a given point x = a, we can associate the power series 

00 

X) a»(x - a) n 

n«*0 

with a n = f {n) (a)/nl which usually converges to /(x) in some interval 
about the point x = a. However, the existence of infinitely many deriva- 
tives at x = a does not ensure the 
convergence of the series Xa n (x — a) n 
to /(x). To ensure convergence, the 
remainder in the Taylor formula (9-1) 
of Chap. 2 must approach zero. 

Inasmuch as every function J{z) 
which is analytic at z = a has in- 
finitely many derivatives at that 
point, we can write down the series 

* f (n) (a) 

£ — 12 (,-«)* 

n« 0 

which converges in some circular 
region \z — a| < r. The question is: 
Does such a series invariably converge to /(z)? 

We prove next (in contradistinction to the situation with the correspond- 

1 See See. 8, Chap. 2. 




563 


SBC. 10 } ANALYTIC ASPECTS 

mg real aeries) that analytic functions can always be represented by power 
series. 

Let f(z) be analytic in some region fi, and let C be a circle lying wholly 
in R and having its center at a. If z is any point interior to C (Fig. 14), 
then it follows from Cauchy's integral formula that 


/(*) 


2iri f — 2 

iUr^fn T ( .-l)/ (r -J ^ (1W) 


But by long division 
1 


i -M + r + •• •+r~ I + 


t n 


i - 1 i - 1 

and substituting this expression with t = (z — a)/(f — a) in (10-1) leads to 

/(z) = — /. ; df + (* - a) I — <# + ••• 

2m L'Cf - a J c (f - a) 2 

(* - a ) n r /(f) 

of- 


where 


fin 


Sc 


/(f) 


2« ^ ( f - a )"(f - 2) 


+ fin. 


Making use of (8-10) gives 

/(z) “ /(<*) +/'(«) (z ~ a) + 


/"(«) 

2! 


(z - a) 2 

/("-!)(„) 

•+ 7 (z - a) n_1 + fin- (10-2) 

(n - 1)! 

By taking n sufficiently large, the modulus of fi n may be made as small 
as desired. In order to show this, let the maximum value of |/(f) | on C 
be M, the radius of the circle C be r, and the modulus of z — a be p. Then 
If — z|>r — p, as shown in Fig. 14, and 

f—M — «| 

JC (!■ - „W- ' 5 1 


|fi»l 


2* 


Jc (f ~ °) n (f ~ z) 


p B M2wr 


Mr /p\" 
r — p \rj 


2x r"(r - p) r - p 
Since p/r < 1, it follows that lim | fi n | = 0 for every z interior to C. Thus, 



564 COMPLEX VARIABLE {CHAP. 7 

one con write the infinite Beries 

f ff (a) f (a) 

m - /(«) + /'(•)(# - a) +'-~ {z - a) 2 + • • • +'— — (z - o)" + • • • 

2! m 

(HW) 

which converges to f(z) at every point z interior to the circle \z — a\ * r. 
The series (XO-3) is the Taylor series of /(z) expanded about the point 
* « a. As in Chap. 2, Sec. 9, one can prove that the representation (10-3) 
is unique. 

Let z a* zo be the singular point of f(z) nearest z - a; then/( 2 ) is analytic 
in the circular region \z — a \ < r 0 , where r 0 ~ \z {) — a\. This circular 

region will then be the circle of 
convergence of the series (10-3) in- 
asmuch as the series diverges for 
1 2 — a | > tq. It should be noted, 
however, that there may be points 
of the region R where f(z) is analy- 
tic which lie outside the circle of 
convergence of this series. How- 
ever, one can always choose a new 
point a about which the expansion 
is performed so that the circle of 
convergence of Taylor’s series about 
that particular point contains with- 
in it the desired value of z as long 
as }{z ) is analytic at z. In this 
manner the region R can be covered 
Fia, 15 by a set of overlapping circles each 

of which is associated with some 

T&ylor-series representation of f(z). 

For example, if f(z) * 1/(1 — z), then the expansion of }{z) about 2 ** 0 yields 
/(*) *1+2 + 2*+.... 

The circle of convergence of this series is \z \ ** 1 . But f(z) ** 1/(1 — z) is analytic at 
f * (%)% (Fig. 15), which lies outside the circle \z\ * 1. If we take a «* t, the formula 
(10*3) yields the series whose circle of convergence is |z — t| « \/5, and this circle in- 
cludes the point z - (J£)t. The reader may find it instructive to deduce the expansion 
for f(z) ** 1/(1 — z) in powers of z — i and determine the radius of convergence with 
the aid of the ratio test. 

PROBLEMS 

1. Expand /(z) «* 1/(1 - z) in Taylor’s series about (a) z - 0, (5) z «■ -1, (c) z * i, 
and draw the circles of convergence for each of the series. What relation do the radii 
of convergence of these series bear to the distance from the point 2 «« 1 to the point 
about which the series exoansion is obtained? 




566 


SEC, 11} ANALYTIC ASPECTS 

2. Expand /(*) » log z in the Taylor aeries about t ■* X, and determine the radius of 
convergence. 

3. Obtain the Taylor expansion about t » 0 for the following functions, and deter- 
mine the radii of convergence of the resulting series: (a) «*, (6) sins, (c) cos*, 
(d) log (I 4- *), («) cosh *. 

4. Expand /(*) — sinh * in Taylor’s series about the point * «* vt, and determine the 
radius of convergence of the resulting series. 

6. Discuss the validity of the expansion (I 4* z) m m 1 -f mt + [m(m — l)/2!]* s *+■ * * • 
for arbitrary values of m. 

6. Verify the expansions: 

(a) 4-i> + l)(* + l)" for |* + 1 1 < 1, 

* n-0 

» (* _ \\n 

(i b ) e* ** e X i — for 1*1 < °°* 

»«o n! 

11. Laurent’s Expansion. We have just shown that a function f{z) 
which is analytic at a given point a can be represented in the neighbor- 
hood of that point in a power series. Moreover, this series represents 
j(z) in the interior of the circular region centered at a and whose radius is 
equal to the distance of a from the nearest singular point f(z). In this 
section we prove a more general theorem due to Laurent. 

Laurent’s Theorem. A function f{z) analytic in the interior and an the 
boundary of the circular ring determined by \z — a\ — R\ and \z — a\ «= R%, 
with R ‘2 < R\ (Fig. 16), can be represented at every interior point of the ring 
in the form 

/(*) = E «•(* - a) n + 2 ~ n ’ (11-1) 

n-0 n~l (« - a)" 

where a n = — <fi df, n « 0, 1, 2, (11-2) 

(f — a) n ^ 1 



Fig. 16 



566 


COMPLEX VARIABLE 


[chap. 7 


L 4 

2m JCt 


m 


2iri r Ct (f - a )-~n-M 


d[ f n * 1, 2,..., 


(11-3) 


Ci and C 2 being the boundaries of the ring. 

To prove the theorem we recall that Cauchy’s formula (8-8), when 
applied to the circular ring, enables us to write 


m 


1 

2 ri 



f — z 2 ri'Ct f — z 


(11-4) 


where 2 is any point in the interior of the ring. 

We show next that the integrals in the right-hand member of (11-4) can 
be represented by the series appearing in (11-1). We begin with the 
integral over Ci and note that if £ is on G 1 and z is in the ring, then 


1 1 1 1 " (z - a) n 

£ - z f - a 1 - (z ~ a)/(£ - a) £ - a (£ - a) n 
since 1 |z — a|/|f — a| < 1. Thus, 


and hence 


J / /(£) dr 


27rt 'Ci £ - 2 


df 


^ (z~a) n 

n-0 (T - a) n+i 

y fjm - °r 
2w®»„tj (f — a) B+1 


dr. 


(11-5) 


Since integration of the series term by term can be justified as in the dis- 
cussion of (10-1), we can write 


_1_I m) dr 


dr 


- £ (* - «)»$ — 
Trt - r ci rr - 


/(r) 




C. (f - a) 


-»+i 


dr 


00 


E a„(? - a)", 
0 


where we define a n hy the formula (11-2). This establishes the equality of 
the first terms in the right-hand members of (11-1) and (11-4). 

We consider next the second integral in (11-4). If f is on C 2f then 


1 


1 


f — z z — a 1 — (£ — a)/ (z — a) 
since |f — a|/|z — a\ < 1 in this case. Hence 


- E 


(r - <*)" 

- n^ n+1 ’ 


n— 0 (2 u) 1 




m 


dr - 


1 


00 

<k E 


/(r)(r - o>» 


‘Note 


2«7c, f - 2 ' 2« ' c« (z - a) n+1 

th “ t r~i“ ) 5 , * iJI 1,1 < h 


dr 



? 


567 


SBC. 11] ANALYTIC ASPECTS 

and the integration of the series term by term now yields 


1 


/(f) 


2 ri fCt f — 2 


df 




2wi n Zi> (z - a) n+1 

o_, 


- E 


where we set 


Q-n = 


n-I (2 ~ a) 

/(f) 


-U 

2x1* JCt (f - a)~ n+1 


df, n » 1, 2, .... 


This establishes the equality of the second terms in the right-hand members 
of (11-1) and (11-4), and the theorem is proved. 

We note that if f(z) is also analytic in the interior of the circle C 2 , then 
the integrand in (11-3) is an analytic function and hence a_ n = 0 by 
Cauchy’s integral theorem. In this case (11-1) reduces to the Taylor 
series, since 

At) ^ f (n) (a) 

2ri f Ci (f — a) n+1 n\ 

by (8-10). 

We can write the series (11-1) more compactly as 


a n = /-.<£ 


f(z) = £ a n(z - 


( 11 - 6 ) 


where the a n can be computed from the formula 

n -°’ ±1 ’ ±2 '- (U - 7) 

and r is any simple closed path 1 which lies in the ring and encloses C 2 . 
It is possible to prove that the representation of f{z) in a given circular 
ring in the series (11-6) is unique. 2 Hence if one obtains for f(z) a repre- 
sentation 

/(*)- E K(z - aT 


in a certain ring with the center at a, the coefficients b n in this representa- 
tion must be identical with those given by formula (11-7). This frequently 
enables one to deduce the Laurent series without evaluating the integrals 
(11-7). 

* Recall that the integrals (11-2) and (11-3) have the same values when calculated 
over any path r into which Ci and C% may be deformed without leaving the ring. 

* See, for example, E. C. Titchmarsh, “The Theory of Functions/' p. 101, 2d ©d., 
Oxford University Press, London, 1939. 



668 COMPLEX VARIABLE [CHAP. 7 

For example, let /(*) — «•/**» and let it be required to obtain the expansion 



Since e* ■■ 1 4 * 4 (f*/2I) 4 * • • 4 (s n /n!) 4 * * • , we have for any * s* 0 


* 

z* 




4 — 4 



This is a Laurent expansion about the origin; hence it is the Laurent expansion about 
the origin. 

The Laurent expansion for e llg , valid for all \z | > 0, can be obtained from the series 

e* *■ 1 4 u 4 (m 8 /2!) H by letting u - l/z. 

As another illustration, consider 


/(*) 


t 

(* — 1)(* — 3) 


( 11 - 8 ) 


This function has two singular points: t - 1 and z=>3. To obtain the Laurent series 
00 

2 dn(z — 1)** valid in the neighborhood of z » 1, we can proceed as follows. Set 
<f>(z) m %/{z ~ 3), and expand <*>(*) in Taylor's series about z » 1. The result is 
z 


3 


1 3 f (j zUl 

2 Ai 2 n+1 


(n-9) 


Since * «■ 3 is a singular point of <*>(*), we conclude that (11-9) converges as long as 
\z — 1 1 < 2. On multiplying this series by l/(z - 1), we get 


» 1 f (* - D w ~ l 

(a - 1)(* - 3) ~ 2(z - 1) »-i 2 n + l ’ 


which is valid for 0 < \z — 1 1 < 2. 

To obtain the expansion of /(*) in (11-8) about z « 3, we set <p(z) «•> z/(z — 1), expand 
it in Taylor's series about z — 3, and multiply the result by l/(z — 3). 

The expansion for /(z) in (11-8) valid for |z| > 3 can be deduced as follows: We de- 
compose /(«) into partial fractions and find 


* « lH _L M 

(« - l)(t - 3) " z - 1 + z - 3* 


( 11 - 10 ) 


But 

1 

1 1 

_i| 

(*+; 

+ i+~) 

for \z \ > 1 

*-l * 

' * i - a/*) 


and 

1 

\ i 

-i| 

(<+; 

+5+-) 

| for |z| > 3. 

* — 3 

■ * 1 - (3/*) 



Substitution of these series in (11-10) yields the desired expansion. 

The reader may find it instructive to obtain the same expansion by writing 

m " (* - 1)(* - 3) " ~t 1 - (1/.) 1 - (3/*) (U41) 

and forming the product of the appropriate series for the factors in the right-hand 
member of (11-11). 



SEC. 12] 


ANALYTIC ASPECTS 


669 


PROBLEMS 


1. Obtain Laurent’s expansions for /(f) *» 1/[*(1 — *)*): (a) about f *• 0, (6) about 
s « 1. 

2 . Obtain Laurent's expansion for e~ 1/,s valid for \z\ > 0. 

8. Expand in Laurent's series about z «* 1: (a) (x — I)®, ( b ) l/(x — 1)* # (c) (s — l) 1 *f 
»/(i - i)*l. 

A Obtain Laurent’s expansion for /(f) « l/((f — 1)(* — 2)) valid in the following 
regions: (a) |s — 1 1 < 1, (fr) |f | > 2, (c) 1 < \z | < 2. Note that in (6) and ( c ) the de- 

oo 

sired expansions have the forms 2 &»*** Show that 


m m 




and 


1 m ly l 
z - 1 “Intif" 


for |f| > 1, 




* — 2 2 \ 2 , 

#. Show that /(z) - 1 /[**(! 


— z)] has the following expansions: 


(o) 53 * n ~ 2 , valid for 0 < \z | < 1, 
0 

(f>) £ va'id for 1*1 > 1. 

n— 0 f 


for |f|> 2. 


12* Singular Points, Residues. If z = a is a singular point of an 
analytic function f(z) and the neighborhood of z = a contains no other 
singular points of f(z), the singularity at z = a is said to be isolated . 

Thus, f(z) = l/z has an isolated singular point z — 0 because the region 
| z | = p > 0 contains no singular points other than 2 = 0 within it. The 
function 


/(*) 


z — 1 
z(z 2 + 1) 


has three isolated singular points: z = 0, z » t, z « — i. The function 

/(*) * 


has two isolated singular points: 2=1 and z — — 1. Not all singular 
points of analytic functions are isolated, however. For example, 


m 


i 

sin (I/ 2 ) 


( 12 - 1 ) 


has a singularity whenever 2 = db(l//cir), k = 1, 2, — These singular 
points are isolated. But (12-1) also has a singular point 2 = 0, which is 
not isolated, for, no matter how small the radius p of the circle ] z\ « p may 
be, this circle contains infinitely many singular points z « db(l/feir) in its 
interior. 



570 COMPLEX VARIABLE [CHAP. 7 

The function log* has a singularity at * » 0, and so does V*. These 
singularities are not isolated because every circle \z \ * p includes part 
of the positive real axis, upon crossing which the single-valued branches of 
log * and \Tz suffer discontinuities if the real axis is chosen to be the cut, 
as in Secs. 16 and 17. The points at which the branches of a multiple- 
valued function assume equal values are called the branch points. For the 
present we shall restrict our considerations to single-valued functions. 

If t m a is an isolated singular point of /(*), then in the neighborhood 
of * — a the function f(z) can be represented by the Laurent series 


/(*) -£«.(*- a)" + Z (12-2) 

»-0 n-1 (Z ~ <*) 

Some coefficients in (12-2) may vanish, and there are two nontrivial cases 
that present themselves: 

1. The expansion (12-2) contains at most a finite number m of terms 
with negative powers of z — a, so that (12-2) reads 


/(*) 


06 


Z o»(2 - a)* + 

»— 0 



a-2 


(z - a) 2 


••• + 


O— «> 

V-ar' 


(1M) 


2. The expansion (12-2) contains infinitely many terms with negative 
powers of * — a. 

The type of singularity at * * a characterized by the representation 
(12-3) is called a pole of order m. A pole of order 1 is also called a simple 
pole. When the expansion (12-2) has infinitely many terms with negative 
powers of z ~ a, the point z = a is called an essential singular point of /(*). 
We shall see in Sec. 14 that the behavior of a function in the neighborhood 
of a pole differs radically from that at an essential singular point. 

We note from (12-3) that whenever f(z) has a pole of order m, one can 
define a function 1 

*00 - (* - a) m f(z), z * a, 

4>(a) » a_ m , 


which is analytic at * «= a, but the function (z — a) m ~~ l f(z) is not analytic 
at * « a. This property is used sometimes to define a pole of order m. 

The coefficient a_i in the Laurent representation (12-2) of f(z) in the 
neighborhood of an isolated singular point * * a plays an important role 
in the evaluation of integrals of analytic functions. This coefficient is 
called the residue of f(z) at z = a. 

When the singularity at z • a is a pole of order m, the residue at a can 

1 When t m a, the function <£{*) assumes the indetenninate form 0/0. We agree to 
define *(a) * lxm <K*). 



t 

BBC. 12] ANALTTIC ASPECTS 571 

be determined without deducing the Laurent expansion. Thus, on multi- 
plying (12-3) by (z — a)"*, we get 

*(*) » (2 - o) m /(z) 

- o_m + 0-_m.fi (2 - a) H + o_i (2 - a)™ -1 + ao(z - a) m -| 


(12-4) 


where a_ m p* 0. Since this is a power-series representation of $(z), the 
coefficient o_i in it must be the coefficient of the term (z — a)”* - " 1 in the 
Taylor expansion of 4>{z) about z = a. Thus 


1 tr- l [(z - o)"/(z)] 
(m - 1)! dz”*- 1 


(12-5) 


We formulate this result as a useful theorem : 

Theorem. If <t>(z) = (z — a) m f(z) is analytic at z = a and <t>(a) p* 0, 
then f{z) has a pole of order maiz — a with the residue given by (12-5). 

As a special case of this theorem we note that when the pole at z = a 
is simple, the residue at a is given by the formula 


a_i = lim /(z)(z — a). 


( 12 - 6 ) 


Example 1. Obtain the residues at the singular points of /(z) — (1 + *)/[*(2 — *)]. 
This function has a simple pole at t — 0 inasmuch as 


*00 


1 + 2 
* z( 2 — z) 


L±i 

2—2 


is analytic and does not vanish at z * 0. Also 


<*>(*) 


(2 - 


2 ) 


1 +* 
2(2 - 2 ) 


L±_i 

2 


is analytic at z * 2 and does not vanish for 2*2. Hence /(«) also has a simple pole 
at 2 * 2. 

The residues at these points can therefore be computed with the aid of the formula 
(12-6). We find that the residue at z * 0 is H end at z * 2 it is — 

Example 2. The function 

i*/ N ^ e$ 

m - ( , + 1 - {z + i){z _ i} 


obviously has simple poles at z 

a~i * lim (z — i) 


— » and 2 * t. Therefore the residue at z * t is 

e* t z t l 

* lim — — * — • 

(z -f i)(z — t) k~+ %z ~ f i 2t 


Similarly, the residue at z * — t is found to be —«~*/2 j. 

Example 3. The function /( 2 ) * l /[*(2 -f 1)*] has a simple pole at z *• 0, since 


*(*) 


1 1 
**<* + 1)« “ (* + l)* 



572 COMPLEX VABUBUB 

is stwiytkf at * - 0 and $(0) i* 0. Therefore, the residue at s — 0 is 


[chap. 7 


The singularity of /(s) at t 


a_i «■ lim 


; - 1 . 


+ 1 )* 

*■ — 1 k a pole of order 2, since 


<Kz) « (* + l ) 2 


is analytic at t -» — 1 and *(~1) * — 1. 
a m -*1 with the aid of (12-5). We get 


<*-1 


i±(i) 

1 ! dx V */*— 1 


1 1 
*(1 -f *)* x 

We can therefore compute 



the residue at 


Example 4. The function (sin r)/* 4 has a pole of order 3 at z * 0 as the reader can 
easily check with the aid of the theorem of this section. Hence the residue at z » 0 
can be computed by using formula (12-5). It is simpler, however, in this case, to writ© 
out the Laurent expansion in the neighborhood of t - 0 and obtain the residue from it. 

Since sin t « z — («*/31) -{- (*V5!) , 


z 4 ■"*» 31* 5! 


for |*| > 0. 


It is dear from this that the singularity at * - 0 is a pole of order 3 with the residue 


-1/31. 

Example 5. The function 

w- coB rh 


has an isolated singular point at * — 
that 


cos u 


1, This point, however, is not a pole, for on noting 


u 2 u 4 
1 ~2! + 4l 


we conclude by the substitution u — l/(z — 1) that for |* - 1 1 > 0, 

1 , 1 1 

008 2 — 1 ~ 2!(z - l) 2 + 4!(z - l) 4 

This is the desired Laurent expansion about * ■* 1. Since it has infinitely many negative 
powers of t — 1, the point z «• 1 is an essential singular point. Inasmuch as the term 
(z — l)" 1 does not appear in the expansion, the residue o_i at z » 1 is zero. 


PROBLEMS 


1. Obtain the Laurent expansions in the neighborhood of the singular points of the 
following functions, and thus obtain the residues: 


(«) , (JO «->'•*, («) j-z - t . C<0 - <•> 9 * </> A 1 ", ( b ) 

W j-Lj . » (W 004 *. w *w *• 


1 -«*• 
** 


,w 


«* 

(*-!)*’ 



sue. 13] ANALYTIC ASPECTS 573 

2. Whenever possible, determine the residues at the poles of the functions in Prob. I 
by means of formula (12-5). 

3 . Obtain the residues in Examples 1, 2, and 3 of this section by deducing appropriate 
Laurent’s series. 

4 . Prove the following theorem: If /(*) » g(z)/h(z) is the quotient of two functions 
analytic at t - a such that g(a) * 0, h(a) » 0, and h'(a) ^ 0, then /(*) has a simple 
pole at * » a with the residue g(a)/h'(a). Hint: Examine the quotient of the Taylor 
expansions of g(z) and h(z) about t •* o. 

5. Use the theorem of Prob. 4 to show that /(*) - cot * « cos s/sin t has simple 
poles at * ■» ±hir, k ■■ 0, 1, 2, .... 

6. Note that f(i) • 1/(2 - i) + l/(* - 1) has the Laurent expansion 

« 1 «0 1 

/(*) - z 5^1*" + E ~ 

£ j > 2 " +l „_1 *" 

valid in the ring 1 < |i| < 2. This expansion has the term 1 /*. Does it follow that 
* - 0 is a singular point of /(*) with the residue equal to 1? 

13. Residue Theorem. Let f(z) be analytic in the given closed region 
R bounded by C, except at the isolated singular points z » z\, z «* z 2l 
Zb,. If these points zt are enclosed by circles F* {k = 1, 2, . . . , m ), 
so that /(z) is analytic in the multiply connected region bounded by C and 
the Tjt, we know that 

<f c f{z) dz =» <f r M dz + ^ rj /(z) dz -| f- (^/(z) dz. (13-1) 

But from (11-7), on setting n - -1, we see that 

(a^) k ~d-6 f(z)dz (13-2) 

2irt •' r* 

where (a_i)t is the residue of /(z) at z = z*. We can thus write (13-1) 
in the form 

m 

dz ~ 2*1 YL (<*-i )*• (13-3) 

,/c Jt-i 


The result embodied in this formula is known as the Residue Theokem: 
The integral of /(«) over a contour C containing within it only isolated singular 
points of f(z) is equal to 2ri times the sum of the residues at these points . 

Inasmuch as the residues of /(«), as demonstrated in the preceding section, 
can often be easily calculated, we see that formula (13-3) provides a simple 
means for evaluating integrals of analytic functions with isolated singu- 
larities. 


Example 1. 


Evaluate 


f 1±L 

Jc *(2 — *) 


dz, 


where C is the circle 


1 . 


The only singular point of the integrand enclosed by C is z - 0. In Example 1 of 
Sec. 12 we saw that the residue of the integrand at i * 0 is Hence the value of the 
integral is (2ri) H «■ ri. The value of this integral over any path C enclosing z m 0 
and f * 2 is 2* i(H - %) « -2vt, since the residues at these points are H and — 



574 


COMPLEX VARIABLE 


[CHAP. 7 


J r «* 

' — — dz over the circular path |*| •> 2. 

C ** + 1 

The residues of the integrand at z - t and z - — i were computed in Example 2, 
Sec. 12. Hence the value of the integral is 

2*i(~ -ir) m 2** sin 1. 

\2 i 2% / 

Example 3. Evaluate j cos <fc. 

We saw in Example 5 of Sec. 12 that z « 1 is an essential singular point with the 
residue zero. Hence the value of the integral is zero for every closed path C which 

does not pass through z m 1. If z - 1 lies on C, the integral is improper and other 

means have to be employed to determine its value. 


PROBLEMS 


1. Use results of Prob. 1, Sec. 12, to obtain values of the following integrals where C 
is the circle |*I *» 2: 

« «’ Lrh- «*> //“'-• « L }=r+ <•> 

* — 2 

2. Determine the residues of f{z) ■» — at z - 0 and z m 1, and thus evaluate 

z{z — 1) 

J r g ~ 2 

— dz, where C is the circle \z\ « 2. 

c *(* - 1) 

f r -f 1 

3. Evaluate the integrals / -r dz (» - 1, 2), where C\ is the circle \z\ «* 1 and 

Ja z ~~ 2z 


Cj is the circle |*| «* 3. 

4. Find the value of / 
Jc 

circle 1*1 - 3. 


i + l 
F C (* — 2)* 


dz, where (a) C is the circle |z| * 1, (6) C is the 


14. Behavior of f(z) at Poles and Essential Singular Points. From 
Laurent’s representation (12-3) of f(z) in the neighborhood of a pole z * a, 
we easily conclude that | f(z) | becomes infinite as z — ► a. The behavior 
of |/(r) | with an essential singularity at z = a is different because the 
expansion (12-2) has infinitely many terms with negative powers of z — a. 
While it is true that in this case | f(z) | as z — ► a is also unbounded, the 
function j f(z) | oscillates as z — ■ » a. Indeed, it was shown by E. Picard 
that in the neighborhood of an essential singular point, f(z) assumes any 
preassigned value, with the possible exception of one value, infinitely many 
times. A discussion of this would carry us too far in the study of analytic 
functions, and we merely illustrate this behavior by an example. Since 


P 


liu 


1 

1 + - + 
Z 



1*1 > 0 , 



GEOMETRIC ASPECTS 


575 


SBC. 15] 

f(z) » e 11 * has an essential singular point at z *® 0. We show that if 4 
is any complex nxunber not zero, there are infinitely many values of z 
in the neighborhood of z « 0 such that 

e l/# * A, (144) 

for on taking the logarithm of (14-1) we get infinitely many solutions 


Log | A | -h i(<f> "b 2kr) 
where 4> is the principal argument of A. 


0, sfcl, ± 2 , 


GEOMETRIC ASPECTS 


15. Geometric Representation. The usefulness of graphical representa- 
tion of real-valued functional relationships in the familiar three-dimensional 
space is too obvious to require emphasis. The customary mode of rep- 
resenting real functions by curves and surfaces fails, however, when one 
encounters functions of more than two independent variables. Thus, a 
relationship u = f{x } y,z) containing three independent real variables x , y } z 
requires a four-dimensionai space for geometric representation. Similar 
difficulties arise when one attempts to represent graphically complex 
functions w = /(z), with z « x + iy. For, to each pair of values (%,y), 
there correspond two values ( u,v ) in w = u + iv, and in order to plot a 
quadruplet of real values {u^v.x.y) we need a four-dimensional space. 

However, a different mode of visualizing the relationship w ® f(z) which 
utilizes two separate complex planes for the representation of z and w is 
possible. The relationship w « f(z) then establishes a connection between 
the points of a given region R in the z plane and another region R ' de- 
termined by w = f(z ) in the w plane. 

On separating w = f(z) into real and imaginary parts one obtains two 
real functions 


u « u{x,y) } 
v = v(x,y), 


(15-1) 


which can be viewed as the equations of a transformation that maps a 
specified set of points in the xy plane into another set of points (u,v) in 
the uv plane. 

We turn now to this mode of studying complex functions. 

Example 1. Let w » z -f a, where a « h -f* ik is a complex constant. 

We aet w *• u + * *• * + iy, and get 

u+w**x+iy + h+ih 

- te 4* W 4- i{y 4* k). 



«76 

Hence 


COMPLEX VARIABLE 


[CHAP. ? 
(15*2) 


u ** % *f k, 

v-y + k. 

Formulae (15-2) are the familiar equations defining a translation, and the relationship 
w — * «f a can be visualised as representing a rigid displacement of points in the * 
plane, where each point is moved h units in the direction of the x axis and k units in the 
direction of the y axis. 

Example 2. To study the function to *• as, where a is a constant, it is convenient to 
use polar coordinates. 

We set x m re m t to « pt**, a - Ae %< * and get 
pe* - Are'< a +*\ 

Hence p ~ Ar, <t> » a + 0. (15-3) 

We see from (15-3) that the modulus of w is got by multiplying the modulus of * by A. 
Also the argument ^ of to is got by adding a constant angle a to the argument 9 of z. 
We can visualise the transformation (15-3) as representing a stretching in the ratio A : 1 
accompanied by a rotation through an angle a. A square with the center at the origin 
in the z plane is thus deformed into a square, a circle of radius R is transformed into a 
circle of radius AR, and more generally any figure is transformed into a similar figure 
enlarged by the factor A, If A ■» 1, we have a pure rotation through an angle a. 

The same conclusions can be reached (but less readily) by setting w « u + iv, z * 
* 4* iy, o •» <*i -f tat and by deducing from w » az the transformation 

u ** a x x ~ 02 y, 

v - Ojx 4 - aiy , 

in cartesian coordinates. 

Example 3. To study the relationship w *■ 1/z, * ^ 0, we again use polar coordinates. 
On setting u> - pe** z m re a , we get pe i<f> « (l/r)e~ a , so that 

1 

P - -» </> « -0. (15-4) 

It is clear from (15-4) that the unit circle |zf ■■ 1 is transformed into the unit circle 
| to | ** 1 in the w plane. Since <t> ■» — 9, the corresponding points on these circles are got 
by reflection in the axis of reals (Fig. 17). As the point A traces out the circle |z| «* 1 in 



Flo. 17 



GEOMETRIC ASPECTS 


t 


sue. 16] 


677 


the clockwise direction, the corresponding point A* in the w plane traces out the circle 
|w| * 1 in the counterclockwise direction. Points in the interior of )*| » 1 are mApped 
into points in the exterior of |tc| * 1, except that the transformation of the point 
2 - 0 is not defined by w « 1/z. Points in the neighborhood of z « 0 map into points 
at a great distance from the origin of the w plane, since p «* 1/r. To complete the 
correspondence of points, we can introduce a new point w * » as the correspondent of 
2*0. The point w * » is called the point at infinity . If we consider the inverse trans- 
formation 2 * l/w, we see that w * 0 corresponds to z * «. 

The reader can show that the equations of transformation defined by w * 1/z in 
cartesian coordinates have the form 


with the inverse 


X 

V 

** + v 2 ' 

~x 2 + „ 4 

u 

V 


y “ u s + «* 


(15-5) 


PROBLEMS 


1. Discuss the transformations defined by (a) w * (1 -f i)z f (b) w « l/(z — 1), 
( c ) w * 1 / 2 , (d) w «* 02 4* 6. 

2. Show that every circle in the 2 plane maps by the transformation w « 1 ft into a 
circle in the w plane if one considers straight lines as the limiting cases of circles. Hint: 
Write the general equation of the circle in cartesian coordinates, and make use of (15-5). 

3 . Show that the bilinear transformation 

02+5 

w wm ; , with ad — be 0. 

C2 + d 

can be decomposed into successive transformations 2 ' * cz -f d, z” «* 1 /z\ w * (o/c) 
+ [(6c — ad)/c]z”, which are the type studied in Examples 1, 2, and 3. Then conclude 
(see Prob. 2) that a bilinear transformation transforms circles into circles. Discuss the 
case when ad — be * 0. 


16. Functions w = z n and z 
determined by the function 


= y/w. Let us study next the mapping 
w * z\ (16-1) 


If we set 2 = re ** and w = pe we get 


so that 


pe** as r 2 e m , 



<f> « 20 . 


(16-2) 


It is clear from (16-2) that the upper half of the z plane maps into the whole 
tv plane, for when s is in the upper half plane, the range of variation of 0 
is 0 < 0 < t. Since <j> = 26, we see that the arguments of the corresponding 
points in the to plane vary from 0 to 2r. Points on the upper half of the 



COMPLEX VAHIA.BXJS 


578 


(chap, 7 


circle \z \ * r map into the entire circle |m? 1 = r 2 (Fig. 18). The half 
ray OA in the z plane maps into the half ray O' A' in the w plane. A 
radial line OB, making an angle 0 with the x axis, goes over into a radial 
line O'B', making an angle <£ = 20 with the u axis. The interior of the 
quadrant OAC of the circle | z | * 1 maps into the interior of the semicircle 
M m 1 in the upper half of the w plane with the boundary ABC going 
over into the boundary A'B'C'. The segment OF of the negative real 
axis in the z plane maps into the segment O'F' along the positive u axis. 
To distinguish points on the positive u axis that correspond to points on 
the ray OA from those on OF, we can imagine that the w plane is slit 



Fig. 18 


along the positive u axis and suppose that the points corresponding to OA 
lie on the upper bank of the slit O' A* and that those corresponding to OF 
lie on the lower bank O'F'. 

The transformation of points determined by (16-1) can be visualized as 
a fanwise stretching of the upper half of the z plane in which the sector 
OAB opens into a sector O'A'B’ and the half circle OACF is deformed into 
the whole circle \w\ = L The semicircles of radius r in the z plane go 
over into full circles of radius p =* r 2 in the w plane. Points in the lower 
half of the circle \z \ = 1 map into the whole circle \w\~ 1, inasmuch as 
the replacement of 0 by d + * in (16-2) yields ~ 20 + 2ir. Thus, two 
distinct points B and G with the arguments 0 and 0 + ir in the z plane 
correspond to one and the same point B' in the w plane. 

This is to be expected, since, on solving (16-1) for z, we get 

z => Vw, (16-3) 

which is a double-valued function. If we set w = pe'* in (16-3), we get 
two values 

2 » \/pe ,< * ,2) , z = Vpe’ 1( * /2) ' Hr! = (16-4) 

For points along the u axis, the argument <f> = 0. Points on the upper 


GEOMETRIC ASPECTS 


SEC. 16] 


579 


bank of the slit O'A' in Fig. 18 correspond to z m Vp, and those of the 
lower bank O'F' to z * — Vp* Thus, along the slit, z » \/w is a dis- 
continuous function unless p =* 0. 

The function 

w «* z n y n a positive integer, (16-5) 

can be studied in the same way. On setting z « re**, te « pe** we find 


p = r w , <p ~ nd. 


(16-6) 


This time a wedge of angle 2 t/r in the 2 plane (Fig, 19) maps into the 



whole of the w plane, and a circular arc ACB of radius R goes over into 
a full circle |ip| = R n . An adjacent wedge OBD of angle 2r/n also maps 
into the whole w plane. If we divide the z plane into a set of n adjoining 
wedges, each of angle 2 ir/n, the entire z plane will be mapped into the w 
plane n times. 

Corresponding to a given point w ^ 0, there will be n values of z de- 
termined by the n roots 


z 




u * 

2 irk\ / 

<t> 2irk\ 

( COS ~ + 

] + i sin I 

~ + ) 

L \ n 

n / \ 

n n / J 


(16-7) 


with &*0, 1, — l. Each of these roots lies in one of the wedges 

into which the z plane is divided. 

Some further insight into the character of mapping by means of (16-1) 
can be gained by studying the maps of lines u = const, v = const. If 
we set z *= x + iy in (16-1), we find 


u ~ x 2 ~ y 2 , 
v = 2 xy } 


(16-8) 


so that the lines u = const, v = const map into orthogonal hyperbolas 
x 2 — y 2 « const, 2 xy = const. Some of these are shown in Fig. 20, in 
which the corresponding points are labeled by like letters. 




Figl 20 


SBC. 17] 

GEOMETRIC ASPECTS 

• 581 

17. Hie Functions w <*» e* and z *= log w. If we set 

u> *» u -f tv sad 

z ** x + iy in 

w « c*, 

(17-1) 

we get 

u + w ** e*~ Hi/ * e*(cos j/ + t sin y)* 


Hence 

u » e* cos y, v « e* sin y. 

(17-2) 


It follows from these equations that 


u 2 + v 2 =» e 2 *, 
v 

- « tan y. 
u 


(17-3) 


Accordingly, the lines x » const map into the circles u 2 + v 2 « const in 
the w plane, and the lines y = const map into the radial lines v/u = const. 
Since 

e '+2kri m e t e 2hri „ ^ fc « Q, dbl, ±2, . . . , (17-4) 


we see that w = c* has an imaginary period 1 2W. Hence, if the z plane is 
divided into horizontal strips of width 2w, with the initial strip determined 
by 0 < y < 2t (Fig. 21), the relations (17-4) ensure that the behavior 



of w * e z in every strip 2kir < y < 2(k + l)ir, k = dbl, db2, . . is 
identical with that in the initial strip. Consequently, we can confine our 
attention to the behavior of w ~ e* in the initial strip 0 <J y < 2ir. 

A segment AC of a straight line x «= x 0 in the initial strip maps by (17-3) 
into a circle u 2 + v 2 *» e 2x °. The points A(x 0r O), C(x 0 ,2t) correspond to 
the same point u * e 2x *, v * 0 on the u axis. The segment OP of the 

1 As for real functions, we say that /(z) is periodic of period a if /(« + o) » /(*). 



COMPLEX VARIABLE 


[CHAP. 7 

y axis maps into the unit circle u 2 + v 2 « 1, since along OP x ** 0; the 
half strip x > 0, 0 < y < 2t, maps into the region |w| >1. If x < 0, 
a segment such &sQR in Fig. 21 maps into a circle whose radius is less than 
1. The half strip x < 0, 0 < y < 2w, goes into the interior of the circle 
\w\ » 1. Points on the lines y ** 0, y = 2ir, forming the boundaries of the 
strip, map into points on the positive u axis. If we slit the w plane along 
the positive u axis, then the points on the upper bank of the slit correspond 
to points on the line p0 and those on the lower bank to points on y * 2ir. 
The interior of the rectangle OACP in Fig. 21 corresponds to the interior 
of the ring between the circles u 2 + v 2 «= 1 and u 2 + v 2 ~ e x °. 

We further note that a point moving along the x axis away from the 
origin O in the positive direction has for its image a point in the w plane 
that moves in the positive direction along the u axis away from the image 
0 ( on the unit circle. A point moving away from O in the direction of the 
negative x axis has for its image a point moving from O' toward the origin 
of the w plane. 

If we consider some definite point w 0 in the w plane, the equation 

w 0 - e a (17-5) 

has for its solution 

z * log Wo = Log | w 0 1 + + 2Jtar), A; « 0, =fcl, =fc2, . . ., (17-6) 

where is the principal argument of w 0 . All these values of z differ only 
by the imaginary part, and therefore there is just one solution of (17-5) 
in each strip 2 kw < y < 2 (k + 1)t. The function 

z = logw 

is therefore infinitely-many-valued. If we restrict our attention to the 
slit w plane so that the argument <t> of w lies between 0 and 2tt, the mapping 
from the w plane to the z plane wiU be single-valued with just one image of 
log w in the fundamental strip 0 < y < 2ir of the z plane. 

To study the map of w = log z we interchange the roles of the z and 
w planes in the foregoing discussion. We remark in conclusion that inas- 
much as all trigonometric functions of z are defined in terms of e *, a study 
of the mapping properties of such functions is reducible to the study of 
mapping by w * e az . 

PROBLEMS 

1. Discuss in detail mapping by the function w « **. 

% Show that the function 

w « a ^2 4- > a > 0, 

maps the circles \z\ » const into confocal ellipses and the radial lines arg z «■ B » const 
into confocal hyperbolas. 



GEOMETRIC ASPECTS 


SEC. 18] 


583 


8. Prove that sin z and tan z are periodic functions. 

4. Show that the curves u(x t y) =* const, v(x,y) ** const in (16-8) intersect at right 
angles (Fig. 20). 


18 . Conformal Maps. We noted in Sec. 15 that the relationship w ** 
f(z) can be viewed as a mapping that sets up a correspondence between the 
points of the z and w planes. If w = fiz) is analytic in some region R 
of the z plane, and if C is a curve in R , there is a remarkable connection 
between C and its image O' in the corresponding region R' in the w plane 
(Fig. 22). Consider a pair of points z and z + Az on C, and let the arc 



length between them be As — PQ. The corresponding points in the 
region R f are denoted by w and w +• Aw, and the arc length between them 
by As' = P'Q\ Since the ratio of the arc lengths has the same limit as 
the ratio of the lengths of the corresponding chords, 


As' I Aw I 

lim — — lim - 

A* —► 0 As Ar -► 0 I Az I 


lim 

Ac — * 0 


Aw ; 


dw 

Az 1 


dz 


(18-1) 


We shall exclude from consideration those points of R at which dw/dz = 0 
because at such points the correspondence of values of z and w ceases to 
be one to one. 1 

Formula (18-1) shows that an element of arc through P, on being trans- 
formed to the w plane, suffers a change in length such that the magnification 
ratio is equal to the modulus of dw/dz at P. This ratio is the same for all 
curves passing through P, but ordinarily it varies from point to point in the 
z plane, since \dw/dz\ need not have the same value at all points of the 
z plane. 

We shall see next that the argument of dw/dz determines the orientation 
of the element of arc As' relative to As. The argument 6 of Az (Fig, 22) 

1 If dw/dz ® f'(z) 0 at some point P of H, then dz/dw «* 1 //'(«) is not defined at the 

corresponding point P' for the inverse function z ~ F{w). Thus F(w) is not analytic at 
P'. Indeed, it can be shown that a necessary and sufficient condition for the existence 
of a unique differentiable solution of w ** f(z) at the point z * zq is precisely f'(zo) p* 0. 



COMPLEX VARIABLE 


584 


{chap. 7 


i a the angle made by the chord PQ with the positive direction of the a 
axis, while the argument 6 l of Aw is the angle made by the corresponding 
chord P f Q f with the u axis. 

Hence, the difference between the angles 6' and 8 is equal to 


arg Aw — arg Az 



since the difference of the arguments of two complex numbers is equal to 
the argument of their quotient. As Az — > 0, the vectors Az and Aw tend 
to coincide with the tangents to C at P and C' at P' y respectively, and 
hence arg dw/dz is the angle of rotation of the element of arc A s' relative 
to As. It follows immediately from this statement that if Ci and C 2 are 
two curves which intersect at P at an angle r (Fig. 23), then the correspond- 



ing curves C\ and C* 2 in the w plane also intersect at an angle r, for the 
tangents to these curves are rotated through the same angle. 

A transformation that preserves angles is called conformal , and thus one 
can state the following theorem : 

Theorem. The mapping performed by an analytic function f(z) is con- 
formal at all points of the z plane where fiz) ^ 0. 

The angle-preserving property of the transformation by analytic func- 
tions has many important physical applications. We shall indicate several 
of these in the remaining sections of this chapter, and we merely note here 
that a number of results deducible analytically from Sec. 15, Chap. 5, 
follow directly from geometric considerations. 

For example, if an incompressible fluid with a velocity potential $(x,y) 
flows over a plane (so that v x = d$/dx, v y — d$/dy) y then it is known 1 
that the streamlines x y y ) — const are (iirected at right angles to the 
equipotential curves $(x,y) *= const. 

1 See Sec. 15, Chap* 5 f and particularly Prob. 6 of that section. 



SEO* 18] GEOMETRIC ASPECTS t 585 

The orthogonality of the curves $ « const and ¥ » const in the z plane 
follows at once from the conformal properties of transformations by 
analytic functions. It was shown 1 that the functions # and W satisfy 
the Cauchy-Riemann equations. One can therefore assert that $ and 
are the real and imaginary parts, respectively, of some analytic function 
w = f(z ) ; that is, 

f(z) = <P(x,y) + i*(x,y). 

But the curves # — const and ^ = const represent a net of orthogonal 
lines (Fig. 24) parallel to the coordinate axes in the w plane, and they are 



transformed by the analytic function w — $(x,y) + iSk(x,y) into a net of 
orthogonal curves in the z plane. 

We saw in Sec. 9 that the real and imaginary parts of every analytic 
function f(z) =* u(x,y) + iv(x,y) are harmonic; that is, they satisfy La- 
place's equation in the region where f(z) is analytic. Since solutions of 
Laplace’s equation are demanded in numerous practical problems, analytic 
functions serve as a useful apparatus for producing such solutions. For 
example, if we take 


w * u + iv =» sin z = sin (x + iy) 
then u + iv « sin x cos iy + cos x sin iy 

» sin x cosh y + i cos x sinh y . 

The harmonic functions u = sin x cosh y f v * cos x sinh y are of special 
interest in deducing solutions of Laplace’s equation in rectangular regions. 2 

Further importance of conformal transformation by analytic functions 
derives from the fact that a harmonic function remains harmonic when 
subjected to such a transformation. If a function <fr(u,v) satisfies Laplace's 


i See Eq. (15-10), Sec. 15, Chap. 5. 

* See, for example, Sec. 20. 


586 

equation 


COMPLEX VARIABLE 


[chap. 7 


d 2 <*> dV 

du 2 ^ dv 2 


0 


(18-2) 


in some region 72' of the uv plane, then <t> still satisfies Laplace’s equation, 
in the appropriate region R of the xy plane, when the variables u, v in 
4>{UjV) are related to x, y by an analytic function 

w = u + tv = f(z). (18-3) 

To see this, construct an analytic function 

F(w) = <t>{u f v) + iypiuyv) (18-4) 

by calculating the conjugate yp{u y v) of the harmonic function 4>(u y v). 

The substitution from (18-3) in (18-4) yields 

F\J{z)) = *{x,y) + i*(x,y), (18-5) 

which is analytic in the region R of the xy plane into which the region 
R f is mapped by (18-3). The function <t>(x,y), being the real part of the 
analytic function F[f(z)] y is harmonic. 

This property of the transformation of harmonic functions by means of 
analytic functions is of the utmost practical importance; for, suppose that 
we are required to find a solution <t>(v y v) of Laplace's equation (18-2) such 
that on the boundary C' of some complicated region R ' in the uv plane, 
4>{u y v) assumes specified values. If it should prove possible to find a func- 
tion w = f(z) which maps the region R' conformally into some simple 
region R (a circle, for example) in the z plane, it may be relatively easy to 
determine the transform 4>(x,j/) of <t>(u y v) in the region R with proper values 
of on the boundary C. 

If $(x, 2 /) is so determined, the function <t>(u y v) can be obtained by re- 
placing the variables in &(x,y) by their values in terms of a and v. It is a 
remarkable fact, first discovered by Riemann, that every simply connected 
region R ' (with more than one boundary point) can be mapped conformally 
onto the unit circle | z | < 1 in such a way that the boundary C corresponds 
to the circular boundary \z \ » 1. 

We shall sketch this mode of solution of the Dirichlet problem in Sec. 21. 


PROBLEMS 

1. Obtain solutions of Laplace’s equation from (a) w « cos z, (b) w » e? f (c) w ■■ 

(d) w m log z, (e) w «* l/z. 

2. Construct the conjugate harmonic functions v(z,y) for the following functions: 
(«) u m cos x cosh y ; (b) u « e* cos y; (c) u » y + e cos y; (d) u » cosh x cos y. 

& Examine the mapping by w ** z 2 and w » z 3 at z 0. Is it conformal at z « 0? 
Examine the behavior of the maps of rays issuing from * » 0. What are the ratios of 
magnification of the arc elements at z 1, z «■ 1 -f i t * m t? 



SBC. 19] 


APPLICATIONS 


587 


? 


APPLICATIONS 


19. Steady Flow of Ideal Fluids. We discussed the flow of nonviscous 
incompressible fluids in Sec. 15 of Chap. 5, where we introduced the con- 
cept of the velocity potential $(x,y) and the stream function ^(x,y). 
These functions were shown to be related by the Cauchy-Riemann equa- 
tions 

34> <H> d * 

dx dy dy dx 

It follows from (19-1) that 

F(z) *= Hz,y) + M(x,y) 


(19-1) 

(19-2) 


is an analytic function of a complex variable z = x + iy. We shall call 
F(z) the complex potential and show that its derivative is related simply to 
the velocity vector v = V$ of the fluid particles. 

By (4-5), 


dF . a * 

dz dx dx 


(19-3) 


and, since v = V4>, so that 

a$> d$ d * 

V x as , V ss xs » 

dx dy dx 

we can write (19-3) in the form 

dF 

— * V x — iv v . (19-4) 

dz 


We shall see in Bee. 21 that because of the simplicity of the complex- 
variable theory in comparison with the theory of real functions, it is often 
simpler to calculate the complex potential F{z) than it is to determine 
either of the real functions $(x,y) or 'f'(x,y). This determination depends 
on certain so-called boundary conditions, which are now to be described. 
We first recall 1 that since v = V4> is orthogonal to the curves $(x,y) = const 
and these curves are orthogonal to the curves x,y ) = const, the vector 
v is tangent to the curves ^(x,y) = const. Hence these curves, called 
streamlines , are the paths of the fluid particles. When a sheet of fluid flows 
past an impenetrable obstacle C (cf. Fig. 25, Sec. 20), the fluid particles 
must flow along the obstacle and hence the boundary C must coincide with 
one of the streamlines. Thus the equation of one of the streamlines, say 

*(x,y) « k, (19-5) 

must coincide with the equation of the boundary C. 

1 See Sec. 3, Chap. 5, and Sec. 18 of this chapter. 



COMPLEX VARIABLE 


688 


[chap. 7 


To determine &(x,y) we must then seek a solution of Laplace's equation 


V 2 ^(x f y) •* 0 


(19-6) 


in the region exterior to the obstacle, which is such that on the boundary C 
takes on a constant value. 

This suggests an indirect mode of solution of the steady-fluid-flow 
problems. One examines the shapes of curves ^(x } y) = const for various 
harmonic functions ^(x,y) } and if a particular curve ^(x } y) = k coincides 
with the boundary C of an obstacle of special technical interest, then the 
function ¥( z,y ) solves a special problem. 

It follows from these remarks that any streamline ^(x,y) = const can 
be regarded as a rigid boundary of some obstacle. 

Instead of determining the stream function we can equally well 

determine a harmonic function $(x,y) which on the boundary C satisfies 
the condition 

d$ 

— - 0, (19-7) 

an 


where n is the unit normal to C, for the statement that the obstacle is 
rigid implies that the normal component v n of v must vanish along C, 
since no particles of fluid can cross C. But v n = n • v, and since v ** V$ 
and 

— « n-V# » v n , 
dn 


we see that (19-7) must hold on C. 

It should be noted that we have assumed in the foregoing that there are 
no sources or sinks in the region and that the fluid is incompressible. 
Moreover, the flow is irrotational, and hence $(x,y) and ^(x^y) are single- 
valued functions. These considerations can be extended to the more 
general situation in which circulation is present. However, as we shall see 
from examples in the following section, the complex potential F(z) will 
then no longer be a single-valued function of z . 


PROBLEM 

Deduce from the boundary condition (19-7) that dV/ds ■* 0 along C, so that ¥ » 
const on C. Hint: Note that d$/dn *• (d$/dx)(dx/dn) 4- (d$/dy)(dy/dri). Make use of 
(19-1), and observe that dx/dn *» dy/d* t dy/dn « -dx/ds on <7. 

20. The Method of Conjugate Functions. We observed in the preceding 
Section that every analytic function F(z) * u(x,y) + iv(x,y) can be 
associated with some flow pattern of an incompressible fluid. In fact, 
every such function determines two flow patterns, since either of the 



APPLICATIONS 


sec. 20] 


589 


harmonic functions u(x y y), v(x,y) can be regards as determining the stream- 
lines. 

The simplest example of an irrotational flow is furnished by the function 


F(z) « cz as $ + i¥ } 


where c is a real constant. Since z » x + iy , we have $ * cx> ¥ ** cy, 
and thus the curves ^ — const are straight lines parallel to the x axis. 
The formula (19-4) for the velocity of the fluid yields v x « c, v v « 0, so 
that the flow is parallel to the x axis. Since div v « 0 and curl v 0, 
there are no sources or sinks in the region and the flow is irrotational. 

As a more interesting example, consider 


-4 + t)- 


4> + i% c > 0, a 2 > 0. 


If we set z * re 1 6 in (20-1), we easily find that 


/ a 2 \ ( a 2 \ . 

& « c I ■ r H 1 cos 0, = c I r 1 sin 0. 


For r » a, we have ^ * 0, and hence the boundary of the circle r » a 
is a streamline. The pattern of streamlines is shown in Fig. 25 by the 
solid lines, and the curves $ « const are indicated by the dashed lines. 





COMPLEX VARIABLE 


590 


{chap. 7 


This flow pattern corresponds to a flow around a circular cylinder, 
velocity components are determined from 


F f (z) 



The 


It is easy to verify that div v = 0 and curl v = 0, so that the flow is 
irrotational. The points for which v x = v v « 0 are z = ±a. These are 
called the stagnation points . 

Let us investigate next the flow pattern determined by 

F(z ) = c log z = u + ivy z = re %6 ) (20-2) 

where c is a real constant. 

If we consider only the one branch of this multiple-valued function for 
which 0 < 6 < 2ir, we get 

F(z) = c(Log r + id), 

so that u = c Log r, v — c6 y 0 < 6 < 2tt. 

If we set ^ « cd, then the streamlines *= const are the radial lines 
and the curves <$> = const are circles c Log r = const (Fig. 26). By Eq. 



(15-1) of Chap. 5, the amount of the fluid crossing per second any closed 
curve C is 


V == J c (v x dy - 



d'f’. 


But ¥ * cd, bo that 

V = c f d6. 

Jc 



APPLICATIONS 


591 


SEC. 20] 


i 


This integral vanishes for any path that does not enclose the origin. If 
the origin z » 0 is within C, then V «= 2vc. Hence for c > 0, the flow is 
outward and we have a source of strength 2ttc at the origin. For c < 0, 
we have a sink of the same strength. Thus, div v « 0 at all points except 
z = 0. 

The circulation J is given by the integral 1 


J °*f c (v*dx + v v dy) = f d* 

and since <t» = c Log r, J = 0 and the flow is irrotational. 

If, however, we take $ * cO and ^ = c Log r, the roles of the curves 
= const and = const in the preceding discussion are interchanged. 
We thus conclude that for this flow the circulation J — 2rc if C encloses 
the origin. This corresponds to the situation described as a point vortex 
at the origin. 

The reader will find it of interest to study the function 

4> + = c ^z H ^ — id log z, a > 0, c > 0, 


for which 'k = const when \z\ = a. The function Sk( x 7 y ) represents a 
flow around a circular cylinder r « a with the circulation 2*rc'. 

As further examples of functions yielding useful solutions of interesting 
physical problems consider the following: 

1. The Transformation w = cosh z. Here 


Thus, 


e* -f- e~ 


w « 

2 


cosh z. 


so that 


or 


u + iv « cosh (x + iy) = cosh x cosh iy + sinh x sinh iy 
*» cosh x cos y + i sinh x sin y , 
u ** cosh x cos y t 
v » sinh x sin y , 


u 


,2 


+ 


cosh 2 x sinh 2 x 
u 2 v 2 


cos 2 y sin 2 y 


1 , 

I. 


1 See Sec. 10, Chap. 5. 



m 


COMPLEX VARIABLE 


[chap* 7 

This transformation is shown in Fig* 27, and it may be used to obtain the 
electrostatic field due to an elliptic cylinder, the electrostatic field due to 
a charged plane from which a strip has been removed, the circulation of 
liquid around an elliptic cylinder, the flow of liquid through a slit in a 
plane, etc. 

The transformation from the » plane to the w plane may be described geo- 
metrically as follows: Consider the horizontal strip of the z plane between 



Pig. 27 


the lines y — 0 and y « t, and think of these lines as being broken and 
pivoted at the points where x » 0. Rotate the strip 90° counterclockwise, 
and at the same time fold each of the broken lines y «* 0 and y =* r back 
on itself, the strip thus being doubly 1 ‘fanned out” so as to cover the 
entire w plane. 

It is interesting to note that this same transformation w = cosh z can 
be used to solve a hydrodynamic problem of a different sort. When liquid 
seeps through a porous soil, it is found that the component in any direction 
of the velocity of the liquid is proportional to the negative pressure gra- 
dient in that same direction. Thus, in a problem of two-dimensional 
flow the velocity components (u,v) are 


u 



v 


k d ± 

dy 


If these values are inserted in the equation of continuity, namely, in the 
equation 

du dv 


the result is 


dx 




d 2 p 

9 ? 


0 . 



APPLICATIONS 


sec. 20] 


/ 


m 


Suppose, then, one considers the problem of the seepage flow under a 
gravity dam which rests on material that permits such seepage. One 
seeks (see Fig. 28) a function p that satisfies Laplace’s equation and that 



satisfies certain boundary conditions on the surface of the ground. That 
is, the pressure must be uniform on the surface of the ground upstream 
from the heel of the dam and zero on the surface of the ground down- 
stream from the toe of the dam. If we choose a system of cartesian co- 
ordinates u } v with origin at the mid-point of the base of the dam (Fig. 
28) and u axis on the surface of the ground, then it is easily checked that 
the function p(u,v) * poy(u f v)/v, where 

w = u + iv = a cosh (x + iy), 


satisfies the demands of the problem. In fact, it was seen in the study of 
the transformation w = cosh z that the line y = v of the z plane folds 
up to produce the portion to the left of u = —1 of the u axis in the w 
plane and the line y ~ 0 of the z plane folds up to produce the portion to 
the right of u » +1 of the u axis. The introduction of the factor a in 
the transformation merely makes the width of the base of the dam 2a 
rather than 2. These remarks show that p(u,v) reduces to the constant 
7T on the surface of the ground upstream from the heel of the dam. If the 
head above the dam is such as to produce a hydrostatic pressure po, one 
merely has to set 

, , vovM 

p(u,v) « 


One can now find the distribution of uplift pressure across the base of the 
dam. In fact, the base of the dam is the representation, in the w plane, 



COMPLEX VARIABLE 


594 


[chap. 7 


of the line x « 0, 0 < y < *, of the xy plane. Hence, on the base of the 
dam the equations 

u — a cosh x cos 


reduce to 


v ~ a sinh x sin y, 

u ~ a cos y, 
v — 0, 


so that p(u,0) = — cos 1 * - • 

x a 

This curve is drawn in the figure. The total uplift force (per foot of 
dam) is 

Po [* a w 

P = — / cos ~ du ~ p 0 a t 

TT J — a (1 

which is what the uplift pressure would be if the entire base of the dam 
were subjected to a head just one-half of the head above tin 4 dam or if 
the pressure decreased uniformly (linearly) from the statu 1 head p 0 at the 
heel to the value zero at the toe. The point of application of the resultant 
uplift is easily calculated to be at a distance b = 3o/4 from the heel of the 
dam. 1 

2. The Transformation w = z 4 e*. One has 
u + iv — x + i y + ( ,x f xv 

= x + iy + e*(eo$ y + i sin y), 
so that u = x + c* cos ?/, 

v — y + e x sin y. 

This transformation is shown in Fig 29 If one considers the portion 
of the z plane between the lines y - ±ir, then the portion of the strip to 
the right of x = —1 is to be “fanned out” by rotating the portion of 
y « +1 (to the right of x ~ — l) counterclockwise and the portion of 
y * —1 (to the right of x = —1) clockwise until each line is folded back 
on itself. This transformation gives the electrostatic field at the edge 
of a parallel-plate condenser, the flow of liquid out of a channel into an 
open sea, etc. 

1 Some material in Secs 18 to 20 is taken by permission from a lecture by Dr. Warren 

Weaver printed in the October, 1032, issue of the American Mathematical Monthly. 



SBC, 21] 


APPLICATIONS 


595 



Fig. 29 


PROBLEMS 

1. Study the flow determined by the complex potential w » cz* in a quadrant x > 0, 
y > 0. The function * «■ 2 cxy can be associated with the flow of fluid around a comer. 

2. Study the flow determined by the complex potential w « c sin % in the semi- 
infinite region |x| < ir/2, y > 0. 

21. The Problem of Dirichlet The procedure for reducing solutions of 
physical problems described in the preceding section is indirect. It depends 
on the examination of various harmonic functions that satisfy the boundary 
conditions appearing in specific physical situations. 

In this section we outline a general procedure for constructing harmonic 
functions which assume preassigned boundary values. Thus, let it be 
required to determine a solution of Laplace’s equation 

V 2 $>(x,y) = 0 (21-1) 

which on the boundary C of a given simply connected region R assumes 
preassigned continuous values 

<i> - (21-2) 

The variable « in (21-2) may be thought to be the arc-parameter 8 measured 
along C from some fixed point. 

The boundary-value problem characterized by Eqs. (21-1) and (21-2) 
is known as the Dirichlet problem, and it can be shown that the solution 
of it exists and is unique whenever the boundary C is sufficiently smooth. 
These conditions are usually met in physical problems. 

We first outline a solution of this problem for the case when the region 
R is the unit circle \z\ < 1 and later indicate how this solution can be 



596 COMPLEX VARIABLE [CHAP. 7 

generalized to yield a solution of the Diriehlet problem for an arbitrary 
simply connected region with the aid of conformal mapping. 

Thus, let it be required to construct in the circle J z j < 1 a harmonic 
function $(x,y) such that on its boundary y (Fig. 30) 

Hx,y) = /(<?), (21-3) 

where f(6) is a specified function of the polar angle 0. 



Instead of determining <t>(x,y), it proves more convenient to determine 
an analytic function 

F(z) « ${x>v) + &(?,V) 9 [*I < 1 (21-4) 

whose real part takes on preassigned values (21-3) and then compute 
$(x,y) by separating F(z) into its real and imaginary parts. Now, since 1 
F(z) + F(z) ~ 2$>(x,y), we can write the boundary condition (21-3) in the 
form 

F{£) + 2/(0) (21-5) 

where f « e %9 represents the values of z — re 10 on the boundary y. If 

1 dr 

We HOW muUipW both members oi (2\-5) by , where z is an 

2iri { - z 

interior point oi the circle, and integrate over y, WB 

1 We use bars to denote the conjugate values, so thatTifij « <*>(*, y) - 
* P rov ® d*&t the conditions (21-5) and (21-6) are equivalent, one must impose 
certain continuity restrictions on /($) usually met in the physical problems, See, for 
example,!. S. Bokoimkoff, “Mathematical Theory of Elasticity,” 2ded n p. 143, McGraw- 
Hill Book Company, Inc., New York, 1956, 



EC. 21] 


applications 


597 


2iri J y f ~ z 2t% h f — z jri *y £ — z 


dt. 


( 21 - 6 ) 


By Cauchy's integral formula, the first integral in the left-hand member 
>f (21-6) is eq ual to F(z). We show next that the second integral has a 
constant value F(0) as long as | z\ < L On expanding F(f) in Madaurin’s 
series, we get 

F(z) - F( 0) + F'(0)z + - F"(0)z 2 +•*•+- F (n) (0 )z n +•■■ (21-7) 
2! n! 


which is convergent for all \z\ < 1, since F{z) is assumed to be analytic 
in \z\ < 1. 

If we set z =* f in (21-7) and form the conjugate F(f), we get 

+ + ^^(o) ?+*•*+ 

2 ! «! 

But on the circle 7 , f = “ \/e* *= 1/f, so that 

W ) + + ( 21 - 8 ) 

£ 2! r n! £ 

The substitution of this series in the numerator of the second integral 
in (21-6) then yields a series of integrals of the form 


1 1 r F n ( 0) 

/ 

n! 2 iri '7 (£ — z)£ n 


0,1, 


But the application of the residue theorem shows that these integrals 
vanish for n > 1, and for n * 0 we get 


2iri J y f — z 


F( 0) = a 0 ~ t&o. 


Thus (21-6) can be written in the form 


F(z) 


i f m 

ri 'y £ — z 


d£ — Oo -f- ib 0, 


(21-9) 


where ao + tho *= F(0). 

The real part oo of F( 0) can be determined explicitly in terms of the pre- 
scribed values f(6) on y, for on setting z *= 0 in (21-9), we get 

i r m 



598 

and therefore 


COMPLEX VARIABLE 


(CHAP. 7 


But f 


Oo =S= 


1 

2 iri 



e % 6 , so that df/f — i dS and hence 


a Q 


1 f2* 

— ) m de. 

2tt j o 


( 21 - 10 ) 


Accordingly, the real part of F(z) is determined uniquely when f(6) is 
known. The real part of F(z) is the desired harmonic function 4 >(j %i/). 
Since f « e l6 } f(0) can be expressed as a function of f, say g( f), and we 
see that the integral in (21-0) has the form 



Integrals of this type can frequently be evaluated in closed form with the 
aid of the theory of residues. 

Formula (21-9) thus solves the general Dirichlet problem for the cir- 
cular region. 

We indicate next how the Dirichlet problem for an arbitrary simply 
connected region R can be solved when the function 

w * w(z) (21-11) 

mapping the region R in the complex w plane conformally onto the circle 

\z \ < 1 is known. Let w — u + iv ; then the desired harmonic function 
&(u,v), assuming the prescribed values 

*(u,v) - *00 (21-12) 

on the boundary C of R } is the real part of some analytic function 

$(w) s v) + (21-13) 

On substituting in 2F(ic) from (21-11), we get 

&M*)] « F{z), 

which is analytic in the circle \z\< 1. 

The values of the real part of F(z) on the boundary y of the unit circle 
are known, since the values of 4>(u,v) on the boundary C are specified by 
(21-12) and the points on C are mapped into points on y by (21-11). We 
can thus write the boundary condition (21-12) in the form 

$ = f{6) on y. 

The substitution of f($) in formula (21-9) then yields F(z)> To obtain 
the desired function 4>(u,c), we must calculate the real part of SF(ic), which 



SEC. 22] APPLICATIONS f 599 

can be determined from F(z) by expressing z in terms of w with the aid 
of (21-11). 

It is clear that the solution of the problem of Dirichlet for an arbitrary 
simply connected domain hinges on the construction of a suitable mapping 
function (21-11). The fact that such a function exists is guaranteed by 
Riemann’s theorem mentioned in the concluding paragraphs of Sec. 18. 
During the past 30 years considerable attention has been given to the 
problem of developing effective methods for constructing conformal maps 
for simply connected domains. 1 A formula for conformal mapping 
of a polygonal region on the unit circle (or alternatively, in the upper 
half of the complex plane) has been supplied 2 by H. A. Schwarz (1843- 
1921) and E. B. Christoffel (1829-1900). 

During recent years extensive applications of complex variables to 
broad classes of problems in the theory of elasticity have been made. 1 


PROBLEMS 


1. Use formula (21-9) to compute harmonic functions «J>(.r y y) in the circular region 

x 2 + ?/ 2 1, which assume on its boundary the following values: (a) 4> « x 2 -j- y 2 t 

(6) <f> ~ x 2 — y 2 , (c) <*> * cos* 0, where 0 is the polar angle. Hint: Note that x x /^{z -{- 2), 
y =» (1/2 0(2 — 2) and that on the boundary of the unit circle 2 = 1 /z. 

2. Set z » m r( cos <t> -f i sin </>), f « e t$ « cos 0 -f i sin 0 in (21-9); take account 
Of (21-10); and stum that the real part <t> of F(z) is 

w - 1 r~-jL=*m* 

2ir Jo 1 — 2r cos (0 — < /») -f r 2 

This formula, giving the values of harmonic function <f> at every interior point ( r,<f > ) of 
the unit circle in terms of the assigned boundary values /(0), is known as Poisson's inte- 
gral formula . (Cf. Chap. 6, Sec. 12.) Because of the difficulty of evaluating real inte- 
grals, this formula is generally less useful than the Schwarz formula (21-9). 

22. Evaluation of Real Integrals by the Residue Theorem. Formula 
(21-9) and the problems in Sec. 21 suggest the use of contour integration of 
complex functions in the calculation of certain real integrals. 

Thus, consider a real integral 


r2r 

I F ( sin 0, cos 0) dO 

J o 


(22-1) 


1 There is a vast literature on this subject, and we cite only a book by L. V. Kantoro- 
vich and V. I. Krylov, “Approximate Methods of Higher Analysis,” Groningen, 1958, 
containing a comprehensive survey of the problem in chap. 5. A useful catalogue of 
mapping functions is contained in the “Dictionary of Conformal Representation,” 
Dover Press, New York, 1952, compiled by H. Kober. 

* This formula is contained in most books on complex- variable theory. See, for exam- 
ple, R. V, Churchill, “Introduction to Complex Variables and Applications,” chap. 10, 
McGraw-Hill Book Company, Inc., New York, 1948. 

1 See Sokolnikoff, op. cit. 



COMPLEX VARIABLE 


600 


[chap, 7 


in which F is the quotient of two polynomials in sin $ and cos 0. The 
evaluation of such integrals, as we shall presently see, can be reduced to 
the calculation of the integral of a rational function of z along the unit 
circle |«| * L Since rational functions have no singularities other than 
poles, the residue theorem (13-3) provides a simple means for evaluating 
integrals of the form (22-1). 

We set z » e 40 , so that dz * e^i dd 


dz 

or dd « (22-2) 

iz 

and we recall Euler's formulas, 

z | z ^ z — z~~~^ 

cos 0 «, , s i n $ = (22-3) 

2 2 i 


On inserting from (22-2) and (22-3) in (22-1) we get the integral 


f c «(*) dz 


(22-4) 


in which R(z) is a rational function of z and C is the circular path \z\ = 1. 
If the sum of the residues of R{z) at the poles within the circle \z\ < 1 
is denoted by Sr, the residue theorem yields j R{z)dz - 2 in Sr, so that 

f2r 

F(mn 0, cos $) dd = 2 riSr. (22-5) 


Example L As a specific illustration of this method of calculating integrals of the 
type (22-1), consider 


/ 





-j- a sin 9 


0 < a < 1. 


(22-6) 


On making substitutions in (22-6) from (22-2) and (22-3), we get the integral 


/ 


f * — - 

Jc tz{ 1 + «(* — z l )/2i] 

2 r dz 

a Jc ** 4" ( 2i/a)z — 1 


where C is the circular path \z\ « 1. 

Since the roots of z 7 4* (2 i/a)z — 1 » 0 are 


z\ 


(1 - VI - a 1 ), 


-1(1 + Vl -e?), 


(22-7) 


( 22 - 8 ) 


2 f dz 

a Jc (z - *l)(z - *») 


we can write (22-7) as 


/ 


(22-9) 



sbc. 22] 


APPLICATIONS 


601 


But it is dear from (22-8) that for 0 < a < 1 we have ]*i| < 1 and |*a| > 1, so that 
only one pole * * z% of the integrand 

B(z) m 1 

W <* - *i)(* - **) 

lies within the unit circle. The residue of R(z) at * » by (12-6), is 

1 


r ■» lim R(»)(i - «i) - 

«-> H *1 — *t 


which, on noting (22-8), yields 


2tVl - a* 


By the residue theorem, the value of (22-9), which is the same as that of the integral 
(22-8), is 

/ - - 2«r - - ^ 

« V l - a 8 

The reader can verify by the same method, or by setting 6 » <p — ir/2, that 

r 2x de _ r 2 r d$ _ 2x 

Jo 1-fa COS ^ Jo 1 + a sin (9 y/l'—aP # 


0 < a < 1. (22-10) 


The infinite integral 

/ °° /(*) 

( 22 - 11 ) 

-00 g{x) 

in which /(x) and g(x) are polynomials in x, can also be evaluated by 
calculating the residues. It should be noted that the integral (22tll) 
converges if, and only if, 1 g(x) = 0 
has no real roots and the degree 
of g(x) is at least two greater than 
that of /(x). 

Now, consider the complex ra- 
tional function 

f(z) 

R(z) - ^ (22-12) 

g(z) 

which, obviously, assumes along the 
real axis the same values as the 
integrand in (22-11). By hypothesis g{z) « 0 has no real roots; hence no 
poles of R{z) lie on the real axis. We form the integral 

f R(z) dz — ( — dz 
Jc Jc g(z) 

where the path C is the boundary of the semicircular region in the upper 
half of the z plane shown in Fig. 31. Since ail roots of g(z) lie at a finite 

1 This follows directly from the usual tests on convergence of improper integrals. 
See Chap. 2, Sec. 8. 




602 


COMPLEX VARIABLE 


(CHAP. 7 

distance from the origin, we can take the radius R of the semicircle Cr 
so great that all poles of R(z) = f(z)/g{z), in the upper half of the z plane, 
lie within the semicircle. If the sum of the residues at these poles is 2r, 
the residue theorem yields 


r /(*) r R fix) r /(*) . v 

/ — » / dx + / dz = 2m£r. 

Jc n(z) J-~Ra(x) ^c R o(z) 


(22-13) 

' g{z) J - R g(x) " ' JcRgiz) 

We show next that when the degree of g(z) is at least 2 greater than 
that of f(z), the integral jT — dz —> Q as R », bo that formula 
(22-13) then yields 

fix) 


r 

J —30 


sO) 


-dx = 27rf2r. 


For proof, set z = in R(z) = f(z)/g(z) t and note that 


m 


0(z) 


M 

W 


M const, 


when R is sufficiently large. Hence, by (5-8) 


I ^ 

1 ■I'—'s 

<f 

M 

1 j Cr g(z) 1 

~ J Cr 

R 2 


M Mir 


(22-14) 


from which it follows that the integral over Or tends to zero as R — ► ». 
Thus under the stated restrictions on f{z) and g{z) ensuring the convergence 
of (22-11), formula (22-14) is true. 

An improper integral like (22-11) should be understood in the sense 


/ Rz r/ r \ 

—4 t/x (22-15; 

g(f) 

Hi -* » 


where R\ and Rt approach infinity in any manner. However, the method of calculation 
indicated in the text actually gives 


lim 
R — > w» 


[ 


1 m 

R <7(^1 


dx 


(22-15o) 


so that «* R 2 in (22-15). The expression (22-3 5a) is termed the Cauchy principal 
value of (22-15). If (22-15) exists (as in the case considered in the text) then obviously 
(22-15a) exists and has the same value. But (22-15 a) may exist when (22-15) does not; 
for example, take f(x) » x, g(x) « 1 + x 2 . 

Example 2. To illustrate the use of formula (22-14) consider an elementary integral, 


L 


« ix 

1 +x 2 


m 


1 _ 1 

l +2* 


Here 



SEC* 22] APPLICATIONS 603 

so that the only singularity of R(z) in the upper half plane is the simple pole at s «* u 
Since the residue of R(z) at * «* i is l/2t, formula (22*14) yields 

r d* . 1 

I ; 9 ** 2lrt -- ** 7T. 

1 + i* 2i 

The essential considerations that have led us to formula (22-14) are: 

1. The integral over the semicircular boundary Cr in (22-13) approaches 
zero as R °o. 

2. The singularities of the integrand in the upper half of the z plane are 
isolated and are at a finite distance from the origin. 

3. There are no singular points on the real axis. 

Clearly, the same procedure can be used to evaluate integrals of the form 


P F(x) dx 

J — -00 

by computing j c F(z) dz as long as the integrand F{z) satisfies conditions 

1, 2, and 3. Occasionally, a slight modification of the procedure 
outlined above can be used when | F(z) | is not sufficiently small in the 
upper half of the z plane, so that the condition 1 is not fulfilled by F(z ) . 
We illustrate this in the following example. 


Example 3. Evaluate 


r 

J — OC 


COS X 

a 2 -f x 2 


dx, 


a > 0. 


If we take F(z ) « (cosz)/(a 2 -f z 2 ), the method outlined above cannot be applied 
directly, since |cosz| = l A\e u 4~ becomes infinite when z -+ » along the y axis. 
However, since cos x is the real part of e**, we can write 



cos x 
a 2 •+• x 2 


dx 



e vx 


dx 


(22-16) 


where Re stands for the “real part of. 0 
Now, if we take 

(22 ' 17) 

then |e w | - | «•(*+*> | » | e ~v| < 1 if y > 0. 
Thus, F(z) in (22-17) is bounded in the 
upper half of the z plane, and there is no 

difficulty in showing that / F(z) dz — ► 0 
Jcr 

as R — ► <». Moreover, F{z) m (22-17) 
has only two singular points, which are 
poles at z\ « ia and zz — ia . Only 
one of these, z\ * ia, lies in the upper half 
plane. Accordingly, 

f - 2 

Jc or + * 



if C is the boundary of the semicircle (Fig. 32) and R is sufficiently large to include the 
point 2 « t\. 



604 


COMPLEX VARIABLE 


(CHAP. 7 

Now the residue of (22-17) at z — ia is e~*/2at t and since j P(t) dx — * 0 as R — * «, 
we conclude from (22-18) that c * 


r e u d* . f e a dx „ . 


j a 2 -f * 2 a 2 + x 2 ' 2ai 

This result is real, and hence the integral in (22-16) is 


„ 

a 


L 


-,dx 


Xt 

a 


* a 2 -f x 2 

Inasmuch as the integrand in (22-19) is even, we conclude that 


J T; 


o a 2 4* x 2 


sdx 


"2a" 


(22-19) 


( 22 - 20 ) 


PROBLEMS 


i: 


f Zv 

1. Use relations (22-2) and (22-3) to write the integrals j j 

2t fa '' 

UV * it e /rtn i i a ii i j * » * t 


d0 


$4 H- sin 0 an< * 

in the form (22-4) and evaluate the resulting integrals by the residue 


'o 6 4- 3 cos 0 
theorem. Check your calculations by formula (22-10). 
r* r de 2x 

% Show that / — -Tj * jrn . 0 < a < 1. 

Jo (14a cos Qj (1 - a 2 )* 

f* x 2 dx r 

3. Show that 

4 . Beferring to Prob. 2, show that 


6* Show that: 


f. 


o (a -b cos 0) 2 (a 2 — 1)^ ’ 

\/2 


if a > 1. 


r° dx 1 r dx tV 2 

* Jo 1 + x 4 2 J->«, 1 -f x 4 4 

m r X 2 dx X 
(6) 

/IT w 

if a > 0. 


r* cos ax 

(e) / — — s dx ~ -e 

Jo 1 


4* x* 

6 . Show that 

7, Show that 


r 


COB xdx X 

iTM 2 ’ a* - ’ 


£ 


dx x (2n — 2)1 

(1 + i 2 ) ft " 2 2 "~* [(n — - 1) !]* ^ 


if n is a positive integer. #*n<; The residue of (1 -J~ z 2 )~ n at * «* « is 

— n(n 4* 1) . • ■ (2n — 2) . 

- 1, 


(n - 1) !2 ,n ~ 1 

One way of seeing this is to let t m z - t, so that (1 4- **)~* - (il)~ ft 2~ fl (l — J#*)'’ 
The coefficient of 1/t is easily found by use of the binomial theorem. 



t 


CHAPTER 8 

PROBABILITY 




Fundamentals of Probability Theory 

1. A Definition of Probability 609 

2. Sample Space 612 

3. The Theorems of Total and Compound Probability 612 

4. Random Variables and Expectation 622 

5 Discrete Distributions 627 

6. Continuous Distributions 631 

Probability and Relative Frequency 

7. Independent Trials 637 

8. An Illustration 641 

9. The Laplace-de Moivre Limit Theorem 6^4 

10. The Law of Large Numbers 650 

Additional Topics in Probability 

11. The Poisson Law 654 

12. The Theory of Errors 658 

13. Variance, Covariance, and Correlation 6611 

14. Arithmetic Means 667 

15. Estimation of the Variance 669 


607 




. . la thSorie des probability riest que le bon sens confirm & 
par le calcul” — Laplace . 


There is no branch of mathematics that is more intimately connected 
with everyday experiences than the theory of probability. Recent de- 
velopments in mathematical physics, moreover, have emphasized the 
importance of this theory in every branch of science. A knowledge of 
probability is required in such diverse fields as quantum mechanics, 
kinetic theory, the design of experiments, and the interpretation of data. 
A recently developed branch of mathematics known as operations analysis 
applies probability methods to questions in traffic control, allocation of 
equipment, and the theory of strategy. Cybernetics, another field of 
recent origin, uses the theory to analyze problems in communication and 
control. In this chapter on probability the reader is introduced to some of 
the ideas that make the subject so useful. 


FUNDAMENTALS OF PROBABILITY THEORY 

1. A Definition of Probability. The idea of chance enters into everyday 
conversation: “It will probably rain tomorrow,” “There may be a letter 
for me at the office,” “I probably won’t get double six on the next throw.” 
It is often possible to assign a numerical measure to the notion of proba- 
bility which these statements illustrate. Such a measure, however, must 
take account of the speaker’s state of knowledge. For instance, in the 
second statement the mailman may know that a letter is there, since he 
put it there himself. His measure of probability and mine are therefore 
not the same. Probability for me is based on my knowledge, and proba- 
bility for him is based on his. 

From this viewpoint (which is one of several possible viewpoints) 
probability is a measure of ignorance. In simple cases the state of ignorance 

609 



610 


PROBABILITY 


[CHAP. 8 

can be accounted for, and probability can be defined as follows: We agree 
to regard two events as equally likely if our ignorance is such that we have 
no reason to expect one rather than the other. For example a 4 or a 6 
is equally likely when a true die is tossed; heads and tails are equally 
likely in a toss of a symmetric coin; aec of hearts and ace of spades are 
equally likely to be drawn from a shuffled deck. 

In the latter example how shall we measure the probability that the 
card drawn will in fact be the ace of hearts? We say that there is “one 
chance out of 52” and define the probability, accordingly, to be If 
it is required only that the card be an ace, common sense suggests that 
the probability should be four times as great, for there are four aces, 
equally likely, and only one ace of hearts. Now, the value %,<i is, indeed, 
the probability that a card drawn at random is an ace. Reasoning in this 
way, we are led to the following definition: 

Definition. Suppose there arc n mutually exclusive, exhaustive , and 
equally likely cases. If m of these are favorable to an event A , then the proba- 
bility of A is m/n . 

The term mutually exclusive means that two cases cannot both happen 
at once; the term exhaustive means that all possible cases are enumerated 
in the n cases. There is seldom difficulty in seeing that these conditions 
are satisfied, but careful analysis is sometimes needed to make sure that 
the cases are equally likely. For example, let two coins be tossed, and 
consider the probability that they both show heads. We might reason 
that the total number of cases is three, namely, two heads, a head and a 
tail, or two tails. Since only one case is favorable, the probability is J/3- 
Now, this reasoning is incorrect. It is true that there are three eases, but 
these cases are not equally likely. The case of a head and a tail is twice 
as likely as the others, since it can be realized with a head on the first 
coin or with a head on the second coin The reader can verify that there 
are four equally likely cases and that the required probability is 

If an event is certain to happen, then its probability is 1 , since all cases 
are favorable. On the other hand if an event is certain not to happen its 
probability is zero, since no case is favorable. By means of the definition 
the reader may also verify the important equation 

g = 1 - p, 

where p is the probability that an event happens and q the probability 
that it fails to happen. 

Since one must begin somewhere, it is impossible to define everything, and every 
mathematical theory contains some undefined terms. These terms should be so simple 
that they are easily understood and also so simple that they are not readily defined in 
terms of anything simpler. The notion “equally likely" is an example of such a term; 
it was explained and illustrated in the foregoing discussion but not defined. 



FUNDAMENTALS OF PROBABILITY THEORY 


SEC. 1] 


611 


Example L If a pair of dice is thrown what is the probability that a total of 8 shows? 
The first die can fall in 6 ways, and for each of these the second can also fall in 6 ways. 
The total number of ways is 


6+6+6+64-6+6 - 6-6 - 36 

and these are equally likely in this problem. A sum of 8 can be obtained in 6 ways, 
namely, as 

2 + 6, 6 + 2, 3 + 5, 5+3, 4+4 
and hence the desired probability is 

This computation of the total number of cases illustrates an important principle of 
combinatory analysis: If one. thing can be done in n different ways and another thing can 
be done in m different ways, then both things can be done together or in succession in mn 
different ways. 

Example 2. In a well-shuffled deck what is the probability that the top 4 cards are, 
respectively, ace, two, three, and four of hearts? 

To find the number of equally likely cases we consider the various possibilities for the 
top 4 cards. The first card may be any one of 52; for each determination of that card 
there remain 51 possibilities for the next; and so on. Repeated use of the principle 
mentioned at the end of the last example gives 

52*51 -50-49 


for the total number of cases. Since only one case is favorable, the desired probability 
is the reciprocal of this. 

When r things are dealt into r numbered spaces from a stack of n distinct things, then 
any particular arrangement of the objects is called “a permutation of n things r at a 
time.” If the total number of such permutations be denoted by n P r , the foregoing 
reasoning yields the important formula 

n P r - n(n - l)(n - 2) . . . (n - r + 1). 

Example 3. If a hand of 4 cards is dealt from a shuffled deck what is the probability 
that the hand consists of ace, two, three, and four of hearts? 

The difference between this example and the preceding is that now the order is not 
relevant. Let C denote the number of distinct 4-card hands, not counting order. Then 
the number of distinct 4-card hands when the order is counted is 

C.4P4, 

since each hand of 4 cards admits 4 P 4 different orderings of its members. On the other 
hand the number of distinct 4~card hands when order is counted is also equal to 
by Example 2. We have, therefore, 

C 4P4 “ &3P4, 

so that 

6sP 4 52-51*50*49 ^ 52! 

C "" 4 P 4 ** 4-3-21 ~4!48l‘ 


The desired probability is the reciprocal. 

When r things are taken from a stack of n things, the groups, so obtained are called 
“combinations of n things r at a time.” If the number of such combinations is denoted 
by n C r , the above reasoning gives the important formula 

^ fJPr n\ 



612 


PROBABILITY 


[CHAP. 8 

In this formula the arrangement of members in a group is not considered. As in the case 
of poker hands, two groups are counted as distinct only if they have different composi- 
tions. 

Example 4. What is the probability of drawing 4 white, 3 black, and 2 red balls from 
an urn containing 10 white, 4 black, and 3 red balls? 

We suppose that the bails are not replaced. The number of ways to get 9 balls from 
the 17 is it C#. The number of ways to get 4 white from the 10 white is 10C4. The 3 
black balls can be chosen in aC$ ways, and the 2 red ones in 3C2 ways. The number of 
favorable cases is found by multiplication (cf. Example 1), so that the desired proba- 
bility is 

10C4 '4C3 'jCj 252 
- 2^3 r 

Example 5. If a number x is chosen at random on the interval 0 < x < 1, what is the 

probability that K < £ < M? 

- f 4 j We imagine the unit interval divided 

into 7 segments each of length Y (Fig. 1). 
* ia * Since the point may be in any one of 

these there are 7 cases, and the phrase 
“at random” ensures that these cases are equally likely. Since only 2 cases are favor- 
able, the desired probability is . 


PROBLEMS 

1. What is the probability that the sum of 7 appears in a single throw with two dice? 
What is the probability of the sum of 1 1 ? Show that 7 is the most probable throw. 

2 . An urn contains 20 balls: 10 white, 7 black, and 3 red. What is the probability 
that a ball drawn at random is red? White? Black? If 2 balls are drawn, what is the 
probability that both are white? If 10 balls are drawn, what is the probability that 5 
are white, 2 black, and 3 red? 

2 . “If 3 coins are tossed, some pair is sure to come down alike. The chance that the 
third coin fell the same way as that pair is Y%\ and hence the probability that ail 3 fall 
alike is What (if anything) is wrong with this argument? What is the proba- 

bility that 3 coins will fall alike? 

4 . What is the probability that a 5-card hand at poker consists of 4 kings and an odd 
card? 5 spades? A sequence in the same suit, such as 2, 3, 4, 5, 6 of hearts? 

6. In how many ways can you seat 8 persons at a table? Arrange 8 children in a ring 
to dance around a Maypole? Make a bracelet of 8 different beads on a loop of string? 

8. The seats in a concert hall are arranged in an ?n by n rectangle, the side m being 
parallel to the stage. What is the chance that a ticket bought at random will be for a 
seat in back? On the side? Somewhere on the outside rows of the rectangle? 

7 . Two dice are tossed, (o) What is the probability that the first die shows 2? (b) Sup- 
pose you are given the additional information that the total shown by both dice is 9. 
What is now the probability that the first die shows 2? (c) If no information is given, 
what is the probability that the total shown is 3? (d) If it is known that the first die 
gave 2, what is now the probability that the total is 3? (Assume that the various 
numbers on the second die are equally likely no matter what is known about the first die.) 

% Sample Space. The equally likely cases associated with the definition 
of probability represent the possible outcomes of an experiment. For 
instance, the 36 equally likely cases associated with a pair of dice are the 
36 ways the dice may fall. Similarly if 3 coins are tossed, there are 8 



f 


SEC. 2] FUNDAMENTALS OF PROBABILITY THEORY 813 

equally likely cases corresponding to the 8 possible outcomes of that 
experiment. The set of all possible outcomes is called a sample space; 
the “points” of the sample space are events. This notion of sample space 
is meaningful even when the events are not equally likely and even if 
there are infinitely many possible outcomes. For technical reasons, how- 
ever, the events composing the sample space are required to be mutually 
exclusive. In tossing a die the events “an even number shows” and “6 
shows” are not suitable for one and the same sample space. 

A finite sample space is one which has only a finite number of points. 
In such a space let the points (that is, events) have respective probabilities 

Pi) VZy • * • ) Vn 

with Pi + p 2 + b pn « 1. 

Suppose the first m sample points, and only those, are favorable to another 
event A. Then we define the probability of A to be 

p(A) * pi + p 2 H h Pm (2-1) 

(and similarly if some other set of sample points is in question). Thus, 
the points of the sample space are weighted according to their probabilities. 

The reader should observe that this definition is consistent with that of 
the foregoing section: If each point of the sample space has the same 
probability 1/n, the result (2-1) becomes 

11 1 m 

P(A) »- + - + •••+-- — 
n n n n 

Sample spaces with constant probability are called uniform. 

For an example of a nonuniform sample space, consider the following 
experiment: Four coins are tossed, and we are interested in the number of 
heads. An appropriate sample space is composed of the events 

no heads, one head, two heads, three heads, four heads 
with respective probabilities, or weights, 

He* He; He; H6; He- 

These values are found by counting cases, as follows. The 4 coins can fall in SS 4 , or 
16, ways. They give no heads in only one case, namely, when they all fall tails, and 
hence the required probability is 3de- To obtain 1 head there are 4 cases: heads on the 
first coin or on the second coin, and so on. This gives }{$• Tor 2 heads, the 2 coins 
giving heads can be any 2 of the 4 coins. Since there are 4 C 2 * 6 ways to choose 2 coins 
out of 4, there are 6 cases favorable to the event, two heads. The probability, then, 
is The other entries are found in the same way, or by symmetry. 

To illustrate the use of this sample space, let us find the probability 
of getting at least two heads. Since the last three points of the sample 



PROBABILITY 


§14 


[chap. 8 


space, and only those, are favorable to this event, the required probability 
is 

He + He + He - l He> 

Again, the probability that there is an odd number of heads is 


He + He * H> 

since that event corresponds to the second and fourth point. On the other 
hand, this sample space does not give the probability that the third coin 
will fall heads, although the underlying uniform space tells us that the 
probability is 

Additional information concerning the experimental situation is apt to 
change the sample space. For example, if a toss of a die is known to have 
given an even number, the probabilities of 1, 3, and 5 are changed from 
H to 0. This question is discussed in Examples 2 and 3. 

When two sample spaces are constructed for a given experiment by the procedure of 
the text, it can be shown that they are consistent; that is, they give the same probability 
for any event to which they both apply. This fact is illustrated in the problems, though 
we do not give a formal proof. 

The notion of sample space enables us to define probability even when there is no 
underlying set of equally likely cases. Suppose we are given n events and a corre- 
sponding set of nonnegative numbers p» such that pi + m H b pn *■ 1. The events 

are said to form a sample space , the numbers p, are called probabilities , and the proba- 
bility of various associated events is defined by addition, as in the text. This abstract 
idea can be extended to sets of very general type, the role of the numbers p, being taken 
by a so-called measure on the sot. With such an approach probability theory is included 
in a branch of mathematics known as the theory of measure. 1 A sample space defined 
with the help of arbitrary numbers p, is considered in Example 1. 

Example 1. A loaded die has probabilities 


PU P*> Ps, Pi, Pb, P« 

of giving the respective values 

1, 2, 3, 4, 5, 6. 

What is the meaning of the condition pi + P 2 H f pe — X? If this condition is satis- 

fied, find the probability that a single toss will give either a 4 or a 6. 

The condition means that one of the stated alternatives will certainly happen; for 
instance, the die does not land on edge. From a more abstract viewpoint, the condition 
means simply that the given events and probabilities form a sample space. When that 
is the case, the probability of getting 4 or 6 is pi -f* p« by definition. 

The assumption that “the probabilities are p % ” is an example of a statistical hypothesis. 
It is an important task of statistical theory to test the validity of such hypotheses by 
examining the consequences. 

The reader should notice that the values pi were not given, and could hardly be 
given, by considering “equally likely cases." They may be estimated, however, by 
repeatedly tossing the die. When p\ is the probability of the ace, it can be shown that 
the proportion of aces actually observed, in a large number of tosses, is likely to be close 

* See Appendix C. 



FUNDAMENTALS OF PROBABILITY THEORY 


SEC. 2] 


615 


to pi* If there are n tosses, and if m aces are observed, this proportion m/n is called 
the relative frequency. The connection between probability and relative frequency is 
discussed in Secs. 8 to 10. 

Example 2. Two coins are tossed. Suppose a reliable witness tells us ,f at least 1 coin 
showed heads." What effect does this have on the uniform sample space? 

The uniform sample space had the following appearance before we received the extra 
information: 


Event 

HH 

TH 

HT 

TT 


Probability 

X 

X 

x 

x 


The new information assures us that the last event is ruled out but gives no indication 
concerning which of the other three may have occurred. Since these three events were 
equally likely to begin with, they arc considered to be equally likely in the new situa- 
tion. (That is not a theorem, but an axiom of probability theory.) The new sample 
space, therefore, is 


Event 

HH 

TH 

HT 

TT 


Probability . . . 

X 

X 

X i 

0 


Example 3. The tossing of 2 coins can be described by the following sample space: 


Event 

no heads 

one head 

two heads 


Probability 

X 

X 

X 


What happens to this sample space if we know that at least 1 coin showed heads but 
have no other special information? 

The first event is ruled out, but we are not told which of the remaining ones occurred. 
It. is an axiom of probability theory that the relative probabilities of the remaining 
events remain unchanged in a situation such as this. Since the event “ 1 head" is twdce 
as likely as “2 heads" in the original Bpace, the same is assumed in the new one. The 
new sample space is therefore 


Event 

no heads 

one head j 

two heads 


Probability 

0 

x i 

X 


(remember that the probabilities must add up to 1). The reader should check that this 
result is consistent with that of Example 2. 

If the events E\, E% ...» Ek of the sample space are the ones favorable to A and have 
probabilities p\ , p%, . . . , Pk, the information that A happened gives a new sample space 
with events E\ } En, . . Ek only. The probabilities on that new sample space are 

cpi, cpz, . . cpk, 

where c is a constant so chosen that the sum is 1 : 

1 

C m * 

Pi + Pi + ' * * + Pk 

This is the general assertion which is illustrated in Examples 2 and 3. 



616 


PROBABILITY 


[CHAP. 8 


PROBLEMS 

1. A coin is tossed 3 times. Construct a uniform sample space for this experiment. 

(That is, make a table showing the 8 possible outcomes HHH, HHT, . . . and their re- 
spective probabilities . , . .) According to your sample space what is the proba- 

bility of at least one H? At most one H? A run of exactly two H’s in succession? A run 
of at least two H’s in succession? H appearing before T? H appearing for the first 
time in the second toss? The sequence THT? The sequence TTT? 

2. In Prob. 1 suppose we are concerned only with the number of H’s. Construct an 
appropriate sample space. (That is, make a table showing the 4 possible outcomes: no 
Ha, one H, . . . with their respective probabilities.) Decide which questions in Prob. 1 
Can be answered on the basis of this sample space, answer them, and verify the agree- 
ment with your answers to Prob. 1. 

3. The following argument is attributed to Leibniz; “A total of 12 with 2 dice is just 
as likely as a total of 11. For, 12 can materialize in just one way, namely, by getting 
6 on one die and 6 on the other; and 11 can also materialize in just one way, namely, by 
getting 6 on one die and 5 on the other.” Using the notion of sample space explain 
what is wrong with Leibniz’ conclusion. (With the uniform sample space, 11 can 
materialize in 2 ways. On the other hand if we choose a sample space in which the 
event “6 on one die and 5 on the other” is a single point, the weight of this point is 
different from that of the point ”6 on one die and 6 on the other.” The student should 
verify these remarks in detail) 

4 . The following is due to d’Alembert: “If we want to get at least one head with 2 
tosses of a coin, heads on the first toss makes the second toss unnecessary. Bo there are 
3 cases, H, TH, and TT, of which 2 are favorable to heads. Hence the probability of 
heads is Discuss, with reference to the uniform sample space and also with refer- 
ence to the sample space ’which has only the three points H, TH, TT. (Ambiguities 
such as this and the preceding can cause serious errors in practice if the notion of sample 
space is not well understood. In fact, one of the reasons for defining the sample space 
is to avoid this kind of difficulty.) 

5. What happens to the uniform space associated with a pair of dice, if we are told 
that the total shown is 7? 

6. Four coins are tossed. A reliable witness tells us that there are at least as many 
heads as tails. What is the most probable number of heads, and what is its probability? 
Suggestion * Use the sample space given in the text. 

7 . A coin is tossed 3 times. If we know that a sequence of 2 tails in a row did not 
occur, what is the probability that a sequence of 3 heads in a row did occur? Suggestion: 
Use the uniform sample space, 

3. The Theorems of Total and Compound Probability. Statements about 
probability are often given an abbreviated notation. If A and B are 
events, AB means the event “A and B”; that is, AB happens only when 
both A and B happen. For example, if two cards are drawn in succession 
without replacing, suppose A is the event “the first draw gives a king” 
and B is the event “the second draw gives an ace.” Then AB happens 
if we get a king on the first draw followed by ace on the second. 

< It is customary to write p(A) for “the probability of the event A.” 
Ih the foregoing example p(A) = since there are 4 kings among the 
52 cards. If nothing is known about the results of the first draw, then 
p(B) « H% also. 



FUNDAMENTALS OF PROBABILITY THEORY 


61? 


BBC. 31 

To see this, note that the total number of cases is 52*51, since there are 52 ways to 
get the first card and, when that card is chosen, 51 ways remain to get the second. To 
count the cases favorable to B , observe that the ace obtained on the second draw may 
be any one of the 4 aces. For each choice of this ace there remain 51 possibilities for 
the first card. The number of favorable cases, then, is 4*51, and hence 


V(B) 


4 51 4 

52 51 “ 52* 


(3-D 


Sometimes two events A and B are so related that the information that 
A happened changes the probability of B. To deal with this situation it 
is customary to write Pa{B) for "the probability of £, given A” In the 
example cited previously, 

Va(B) = Hi (3-2) 

(for if A happened, the first draw gave a king, and hence the 4 aces are 
to be found among the remaining 51 cards). On the other hand when A 
is the event “the first draw gives an ace” and B t as before, is the event 
“the second draw gives an ace,” then Pa {B) ~ (since now only 3 
aces remain when A happens). Both values for Pa(B) are different from 
p{B), the probability of ace on the second draw when nothing is said about 
the first draw. 

In this notation the theorem of compound probability takes the following 
form : 

Theorem If A and B are any events , then 

p{AB) - p(A)p A (B ). (3-3) 

Informally, “the probability that A and B happen is the probability 
that A happens times the probability that B then happens.” A proof is 
easily given by considering equally likely cases. Let n (l , n bl and n ab denote 
the numbers of cases favorable to A y B, and AB, respectively. Then 


P(AB) = 


riab 


U a Tl a b 
n n a 


Now, n a /n is p(A) by definition. After A has happened, the only possible 
cases are the n a cases favorable to A. Of these, there are n ab cases favor- 
able to B. Since the n a cases are to be considered equally likely, the 
quotient n a i/n« represents the probability of B when it is known that 
,4 happened, and this gives (3-3). 

To illustrate the theorem (3-3), let us find the probability of drawing 
2 aces in succession from a pack of 52 cards. The probability of ace on 
the first trial is %%. After the first ace has been drawn, the probability 
of drawing another ace from the remaining 51 cards is 2£i> so that the 
probability of two aces is 


%2 ‘Hi * /I221- 



PROBABILITY 


018 


[chap. 8 


This assumes that the first card is not replaced. When it is replaced, the 
reader will find that the desired probability is 


%T%2 ** K69- 

For another illustration of the theorem (3-3), lot us find the probability 
of drawing a white and a black ball in succession from an urn containing 
30 black balls and 20 white balls. Here the probability of drawing a white 
ball is 2 %o- After a white ball is drawn, the probability of drawing a 
black ball is 3 %g. Hence the probability of drawing a white ball and a 
black ball in the order stated is 


V « 2 %o-% - l H 9- 

The events A and B are said to be independent if the information that 
A happened does not influence the probability of B . Hence for such 
events Pa(B) « p(B), and the theorem of compound probability takes the 

form 

p(AB) = p(A)p(B) } for independent events. (3-4) 


For instance, let a coin and a die be tossed, and let A be the event “head 
shows’’ while B is the event “4 shows.” These events are independent, 
and hence the probability that heads and 4 both appear is 

p(AB) « p(A)p(B) - (H)Q4) - i{ 2 . 

The result (3-4) is readily extended to any number of independent events 

Besides the theorem of compound probability, there is a second funda- 
mental relationship, known as the 
theorem of total probability. If A 
and B are two events, A + B is 
defined to be the event U A or B 
or both.” For instance, let A be 
the event “a number greater than 
3 shows” while B is the event “an 
even number shows” in a toss of 
a die. Then A + B happens if the die gives 2, 4, 5, or 6. In this 
notation the theorem of total probability reads as follows: 

Theorem. When A and B are any events , then 

p(A + B) = p(A) + p(B) - p{AB). (3-5) 

We can represent the statement (3-5) diagrammatieally by the intersecting 
point sets A and B shown in Fig. 2. 

Referring to the definition of probability by equally likely cases, suppose 
the numbers of cases favorable to A, B, AB, and A + B are denoted by 




SEC. 3] 


FUNDAMENTALS OF PROBABILITY THEORY 


619 


tl af Tll» Hab) ^04 

respectively. To find the number favorable to A + B f it will not do 
simply to add n a and rib , for the eases favorable to both A and B are counted 
twice in this addition. To take account of that we must subtract n a i, 
thus: 

n a +b = n a + n b - n ab . 


Dividing by n, the total number of cases, gives 

Wa-f-6 Ra ^ H b Hab 

n n n n 

which is equivalent to (3-5). 

To iliustrale the theorem, let us find the probability that at least one 
die gives 4, when two dice are tossed. The probability that both give 4 
is The probability that the first gives 4 is and similarly for the 

second. Hence the probability that at least one gives 4 is 

p(A + B) =* + 14> — M6 “ l /i$- (3-6) 

Thus is consistent with the icsult given by counting cases Specifically, there are 5 
cases with a 4 on the first and a number other than 4 on the second, there are 5 cases 
with 4 on the second and a number other than 4 on the first, and there is I case with 4 
on both The number of favorable cases is therefore 5+5-fl *11 so that (S-ti) 
follows 

For mutually exclusive events, that is, for events A f B which cannot 
both happen, p(AB) = 0. Hence the theorem of total probability takes 
the form 

p(A + B) — p(A) + p(B ), for mutually exclusive events. (3-7) 

The statement (3-7) can be depicted 
by the nonintersecting point sets in 
Fig 3 

For example, in a toss of a die let 
A be the event “4 shows” while B FlG - 3 

is the event “5 shows,” Since these 

events are mutually exclusive, the probability of getting either 4 or 5 is 

p(A + B) - p(A) + p(B) « H + H = 

A result similar to (3-7) applies to any number of mutually exclusive 
events A, B, ( 7 , 

The foregoing analysis, by counting cases, establishes the theorems of 
total and compound probability for uniform sample spaces only. Actually 
the results are valid for arbitrary sample spaces, as will be indicated next. 




PROBABILITY 


620 


[chap. 8 


Assuming that the sample space is finite, let the events Ei of the sample 
space be so numbered that 

Ei y . . Ej 

are favorable to A alone, 

• ♦ •> &k 

are favorable to both A and B , while 


Eic+i, . . . , E m 

are favorable to B alone. If the associated probabilities are p t , then (3-5) 
is equivalent to the identity 


Pi H h Pm - (Pi H f py + Py+i H bpk) 

+ (pj-fi H b Pfc + Pk+i H h Pm) — (Pj-fi H h p*). 

The three parentheses on the right represent, respectively, p(A), p(B) f 
and p(AB) by definition. 

To derive (3-3) for a general sample space, recall that the sample points 
favorable to B have the same relative weights after A happened as before. 
Hence, in the previous notation, 


p(AB) = p j+ 1 h Pk 


(pH b Pk) 


Pj+i 


<Pi H b Pk 


■+•••+■ 


Pk 


PiH b pk) 


« p(A)p^(^). 

Example 1. The probability that Peter will solve a problem is pi, and the probability 
that Paul will solve it is p%. What is the probability that the problem will be solved if 
Peter and Paul work independently? 

The probability that both solve it is Pipi, by the theorem of compound probability, 
(3-3). Hence the probability that at least one solves it is 


Pi + P2 ~ pm (3-8) 

by the theorem of total probability, (3-5). 

Example 2. Solve Example 1 by finding the probability that both fail. 

Peter's probability to fail is 1 — pi, and Paul’s probability to fail is 1 — p 2 * The 
probability that both fail is 

(1 - p t )(\ - pa) 

and the probability of the contrary event, that at least one succeeds, is 

1 - (1 - Pi)(l - P 2 ). (3-9) 

The consistency of (3-8) and (3-9) is easily verified. 

Example 3, A bag contains 10 white balls and 15 black bails. Two balls are drawn 
in succession. What is the probability that one of them is black and the other is white? 

The mutually exclusive events in this problem are (a) drawing a white ball on the 
first trial and h black ball on the second, (6) drawing a black ball on the first trial and 



SBC. 3] FUNDAMENTALS OF PROBABILITY THEORY 621 

a white on the second. The probability of (o) is • 1 and that of (6) is • ! % 4l 
so that the probability of either (a) or (i>) is 

10 Ai + - M. 

Example 4. How often must a pair of dice be tossed to make it more likely than not 
that double 6 appears at least once? 

The probability that double 6 does not appear on a given toss is s %e>, no matter 
what is known about the preceding tosses. Repeated use of the theorem of compound 
probability gives 

(•He)" 

for the probability that double 6 does not appear in any of n tosses. It is desired to 
choose n in such a way that this probability is less than l A • Thus, 

{*%*)" < K. 

Taking the logarithm gives 

< - log 2 


or 


n > 


log 2 
log 3 %5 


24.6. 


Thus 25 tosses suffice, but 24 do not. 

Example 5. Peter and Paul take turns tossing a pair of dice. The first to get a throw 
of 7 wins. If Peter starts the game, how much better arc his chances of winning than 
Paul’s? 

This problem is different from any we have considered hitherto, in that there are 
infinitely many possibilities. Namely, Peter may win on his first throw, or on his second 
thiow, or on his third throw, and so forth. To apply the preceding theory, we simpty 
consider the probability that Peter wins in n throws and take the limit as n oo. 
A wide variety of questions involving infinitely many outcomes may be dealt with in a 
similar maimer. 

The probability of 7 is 1<>, and the probability of not getting 7 is %. Ifcnce the 
probability that Peter wins on his first throw is The probability that Peter wins 
on his second throw is (?o) 2 (/6) (since Peter’s first throw and Paul s first throw must be 
other than 7 but Peter s second throw must be 7). Peter’s probability of winning on his 
third throw is (^o) 4 (3ti)* and so on 

By the theorem of total probability the probability that Peter wins is 
H + ( H)(K ) 2 + (hXH) 4 + • • * - ( W + r 4- r 2 + . • .), where r * (%)* 


1 l 1 1 6 

“ei-r'ei - * li ' 


(3-10) 


A similar procedure shows that Paul’s chance of winning is V\i, or one can reason as 
follows: The probability that 7 does not occur m n trials is ( 5, g) n . Since the limit is 
zero, the probability of an eternal game is zero, and Peter or Paul is sure to win. Thus, 
Paul’s chance is 

1 - Hi * Hi- 


PROBLEMS 

1. What is the probability that 5 cards dealt from a pack of 52 cards are all of the 
same suit? 



622 PROBABILITY [CHAP. 8 

3 . Five coins are tossed simultaneously. What is the probability that at least one of 
them shows a head? All show heads? 

8 . The probability that Paul w r ill be alive 10 years hence is % and that John will be 
alive is What is the probability that both Paul and John will be dead 10 years henoe? 
Paul alive and John dead? John alive and Paul dead? 

4 . One purse contains 3 silver and 7 gold coins; another purse contains 4 silver and 8 
gold coins. A purse is chosen at random, and a coin is drawn from it. What is the 
probability that it is a gold coin? 

5* Paul and Peter are alternately throwing a pair of dice. The first man to throw a 
doublet is to win. If Paul throws first, what is his chance of winning on his first throw? 
What is the probability that Paul fails and Peter wins on his first throw? 

8. How many times must a die be thrown in order that the probability that the ace 
appear at least once shall be greater than Hs? 

7. Twenty tickets are numbered from 1 to 20, and one of them is drawn at random. 
What is the probability that the number is a multiple of 5 or 7? A multiple of 3 or 5? 

Note that in solving the second part of this problem, it is incorrect to reason as follows: 
The number of tickets bearing numerals that are multiples of 3 is 0, and the number of 
multiples of 5 is 4. Hence the probability that the number drawn is either a multiple 
of 3 or of 5 is %o 4* Ho *= Vi- Why Is this reasoning incorrect? 

8 . A card is chosen at random from each of 5 decks What is the probability that all 
are face cards? Would the probability be larger or smaller if all 5 cards were taken from 
one deck, without replacing? 

9. Answer the two questions in Prob. 8 if the desired hand is 1, 2, 3, 4, 5 of clubs; if 
the desired hand is to have at least 2 aces but is otherwise unrestricted. 

10 . Each of two radio tubes has probability p of burning out during the first 100 hr 
use. If both are put into service at the same tune, what is the probability that at least 
one of them is still good after 100 hr? Generalize to n tubes. If p « 0.1, how many 
tubes are needed to give a probability > 0.99 that at least one is good after 100 hr? 

4. Random Variables and Expectation. 1 A process is random if it is 
impossible to predict the final state from the initial state (as, for example, 
in a toss of a coin or a die). Associated with a random process there may 
be certain numerically valued variables which themselves have a random 
character. For instance, if A" denotes the number obtained by tossing a 
die, then X is a variable which assumes the values 

1, 2, 3, 4, 5, 6 

corresponding to the six events: 1 shows, 2 shows, and so forth. The re- 
spective probabilities are 

H, K, K, H> H> H- 

Again if X is the number of heads obtained when 3 coins are tossed, then 
X is a variable which assumes the values 

0, 1, 1, 1, 2, 2, 2, 3 (4-1) 

Actions 4 through 6 may be omitted on the first reading without loss of continuity, 
but they are essential to the developments in Sec. 13. 



SEC. 4 ] FUNDAMENTALS OF PROBABILITY THEORY 623 

corresponding to the various ways the coins may fall. For instance, 
X *= 2 corresponds to each of the throe events: HHT, HTH, THH. 

Similarly, if a gambler stakes d dollars on a game, the amount he wins 
assumes the values 

d y —d 

in correspondence with the events “he wins the game” and “he loses the 
game.” If his probability of winning is p, the respective probabilities of 
X = d and X = —d are 

p, 1 - p. 

These special cases illustrate the important idea of random variable . A 
random variable is a numerical-valued function defined on a sample space 
In symbols, 

X(c t ) = x t) i = 1, 2, n, (4-2) 

where e x are the events of the sample space and x t are the values of the 
random variable A". 

Let (e t ) be a sample space of n events c, with associated probabilities 
p t . Let X be a random variable defined on jc,) and assuming the value 
x x at the fth sample point, so that (4-2) holds. The expectation or expected 
value E(X) is then defined by 

E(X) = PiJi + P2*2 A f~ Pn*n. (4-6) 

For example, if X is the number obtained in a toss of a die, then X 
assumes the values 1, 2, 3, ... with corresponding probabilities p % = 34* 
I fence 

E(X) = H , l + K2-h34'3 + 3^*4 + K*5 + J^*6 = %. 

Similarly, if A" is the number of heads obtained when 3 coins are tossed, 
then (4-1) and (4-3) give 

E(X) ~ /8 + H + H+M + H + % + % + % - 

when we note that p, — 1 # in this ca.se. 

By grouping terms we can write t he above sum in the form 

E(X) « J^‘0 + V! + V2 + H-3. 

The factors 

0, 1, 2, 3 

represent the numerically distinct values of X, and the factors 

Vs, H, H, H 

represent tlxe probabilities corresponding to these distinct values. For 
example, % is the probability of 2 heads when 3 coins are tossed, and 



m 


PROBABILITY 


{CHAP. 8 

hence % is the probability that X « 2. A similar grouping of terms can 
be applied to the general definition (4-3) and yields the following useful 
theorem: 

Theorem I. The expectation E(X) is given by 

E(X) = P\X\ + P 2 X 2 H — * + P r x r 

where x X) x 2 , . . , , x r are the numerically distinct values of X and where Pi 
is the probability that X = x<. 

Let Xi be the r distinct values of a random variable X, and let yj be the 
9 distinct values of another random variable Y. The sum X -f Y is a 
random variable which is defined to be x % + y 3 when X ~ x, and Y = yj. 
Thus, X + Y is defined on a sample space whose points consist of the rs 
events 

X = Xj and Y ~ Vj (4-4) 

for i « 1, 2, . . r and j = 1, 2, . . s. One of the most important theo- 
rems in probability theory concerns sums of variables and reads as follows: 

Theorem II. The expectation of the sum of two random variables is equal 
to the sum of the expectations, or in symbols } 

E(X + Y) « E(X) + E(Y). (4-5) 

To prove Eq. (4-5) let p tJ be the probability that simultaneously X = x t - 
and Y = yj. Thus, is the probability of the event (4-4). The definition 
of expectation yields 

E(X -f Y) =23 Pa( x i + 2/y), (4-6) 

since Xi + yj is the value of X + Y which corresponds to the event (4-4). 
By rearrangement, 

E{X + Y) = Z (Z P.y) + Z Vi (Z Pa) • (4-7) 

Now, ZP»; j represents the probability of 

(X * x t , Y ** yi) or (X = x if Y = y 2 ) ... or 

(X - Xi, Y « y s ). 


Hence, it represents 1 the probability P % that X = x t . Theorem I now gives 

Z *«• (Z P»y) - Z x t Pi~E(X) 


and similarly, 


Z P; (Z Pa) = E(Y). 


1 This shows that 2 Xpy * 2P, » 1, hence that the events (4-4) actually do formi 
•ample space. 



FUNDAMENTALS OF PROBABILITY THEORY 


SEC. 4] 


625 


Thus, (4-7) is equivalent to (4-5). The extension to any number of vari- 
ables is immediate. 


The following alternative approach to Theorem II does not require the use of Theo- 
rera I. Let X be defined on a sample space {a*} containing n events and Y on a space 
| bj] containing m events. Thus, 

X(<h) ** Zt and Y(bj) - yj. 

The variable A" Y is defined on a sample space whose mn events e tJ happen when, and 
only when, a t and b 3 both happen. The value of X 4* Y corresponding to the event e,y 
is defined to be -+* Vy If P 13 is the piobabiiity of e lJ} then the definition of expecta- 
tion gives (4-(>), which may be written in the form (4-7) as betfore. Since the events of 

the sample space {M are mutually exclusive (Sec. 2), the sum 2D Pv represents the 

J 

probability of 


a, and b\ or o, and 62 ... or a,- and b m . 

Hence it represents p t , the probability of a x . The first term in (4-7) is therefore E(X) 
by (4-3), and similarly, the second term is E(Y). 

The sums 

L, Vtj and 2J P*j (4-8) 

j * 


are called the marginal probabilities of a, and b Jy respectively. In modern statistical 
theory it is customary to start with the larger sample space jc v { and to define the 
probabilities on the smaller spaces |o t | and |fr ; | by means of (4-8). Theorem II is then 
valid, so to say, by fiat. 


Since Xp x « 1, 
center of mass: 


the expectation E( X) in (4-3) may be interpreted as the 


E(X) - 


Pl*\ + p 2 X 2 H h PnXn 


Pi + P2 d h Pn 


For equally likely x t the result reduces to the arithmetic mean 

1 1 

E( X) » - (xi + x 2 H b x n ), if each = — 

n n 

Thus, E(X) is a measure of the location of X; it is a typical value . The 
following sections show that if sufficiently many observations of the 
variable X are made, the mean of those observations will almost certainly 
be close to E(X). In this sense, E{X) represents the average value at- 
tained by X in the long run. 

Throughout this section random variables were denoted by capital letters to avoid 
confusion between the variable and its values x % or y 3 . In statistical literature the varia- 
bles are usually denoted by small letters. Since the distinction has now been sufficiently 
emphasized, we shall often use small letters in the remainder of this chapter. Thus, 
depending on the context, x x may be a set of random variables or the values of a single 
variable. 



626 


PROBABILITY 


[CHAP, 8 


Example 1. Find the expected number of heads when n coins are tossed. 

Let X* *= 1 if the ith coin shows heads and X» » 0 otherwise. Then, for each i t 


E{Xi) « H I 4 - H O m 

(The reader is cautioned that X\, X% ... are distinct variables here, not the different 
values Xi of a single variable.) The number of heads m is 


and hence 


m » X\ + X 2 4 — * 4 - X n} 

E(m) -ElXi+X* + --- + X„) 

= E( Xi) + E(X Z ) + E(X n ) 




n 

2 * 


Example 2. From an urn containing a white and 6 black balls, a ball is drawn at ran- 
dom and set aside. What is the expected number of white balls left in the urn? 

Let X be the number of white balls left. If a white ball is drawn, then X ** a — 1, 
whereas if a black ball is drawn, then X * a. Hence 


E(X) 


a +b 


(«-!) + 


a -f b 


a 

a -f- b 


Example 3. A deck of cards is thoroughly shuffled. We say there is a coincidence if 
a card has the same position after shuffling as it had before (e g , if it is the fourth from 
the top both times). Find the expected number of coincidences. 

Let A r , = 1 if the tth card is in the same position before and after shuffling, and let 
X, ** 0 otherwise. Then 

E(X % ) - J< 2 .1 + 6 « 2 0 - 

Since the number of coincidences is XX its expectation is 

tf(Xi) + E(X t ) + ..-4 -E(Xri - 1. 


PROBLEMS 

1. A bent coin has probability p of giving heads and probability q « 1 — p of giving 
tails. Let X be a random variable representing the number of heads when the coin is 
tossed three times; X is defined on a sample space consisting of the 8 events HHH, 

HHT, . . . with associated probabilities p 3 , p 2 q t (a) Make a table giving the 8 values 

of X associated with the 8 sample points and their respective probabilities, (b) Make 
a second table giving the 4 distinct values of X and their probabilities, (c) Compute 
the expectation E(X) from your table (a) and also from your table (6). 

2. If X is the number of heads and Y the number of tails, find E(XY) from your 
table in Prob. la and also from that in 16. Is it true that E(X Y) * E(X)E(Y)? Hint 
Make a table giving the 4 values of AT in the 4 cases of Prob. 16. 

8. Peter turns up the cards one at a time from a 52-card deck, and Paul tries to guess 
what the cards are. Find the expected number of correct guesses (a) when Paul calls 
out at random, perhaps repeating himself, (6) when Paul calls off the 52 cards, naming 
each one just once, (c) when Paul calls out “ace of spades" each time. (Assume that 
" Paul has no actual insight into the behavior of the cards.) 

4. In Prob. 3, suppose Peter tells Paul what the card was immediately after Paul 
guesses. Paul has ihe good sense not to call any of those cards, since he knows they 



FUNDAMENTALS OF PROBABILITY THEORY 


627 


SEC. 6] 

have been set aside* What is the expected number of correct, guesses now? Hint: Let 
A\ *» 1 if the rth guess is correct, X* *■ 0 otherwise. E(X t ) « ? The expected number 
of correct guesses will t>e found to be approximately log# 52. 

5 . A coin is tossed repeatedly. What is the expected number of the toss at which 
heads first appear? Hint: Let X be the numt>er of the toss at which heads first appear. 

Then X has the values 1, 2, 3, ... with resjxjetive probabilities V% 34 , H, The 

reader is reminded that Xnr n * r/(l — r ) 2 for any r such that |r | < 1. 

5. Discrete Distributions. When the values x x of a random variable are 
distinct, the associated probabilities p t may be written in the form 


P. = /(a-.). 

Since the ,r,' are supposed to be all the possible values of x, we must have 

2/0 r.) = 1, (5-1) 

just as in the last section 2 p t = 1. Also fix) > 0, because fix) is a proba- 
bilitv. 

For example, let .r be the number of heads obtained when 4 coins are 
tossed. If the value x — 0, 1, 2, 3, or 4 is given, then the probability to 
assume that value is determined by the table 


X «= 

0 

! 

1 

2 

3 

4 

fix) - 

1 16 

6 


*16 i 



The function /(x) is called the frequency function for reasons which will 
now be explained. Suppose n observations of the variable x are made; 
how often should we expect .r = x t ? To answer this question, let Xk = 1 
it .r ~ x t at the kth observation and Xk — 0 otherwise. The number of 
times x =• x, is 

m — X i + X<z + * * * + X n . 

Since the definition of expectation gives 

E(X k ) = 1 •/(*,) + 0[1 - fix,)) = f(x t ), 

we have the fundamental result 

E(m) = nfi x t ). (5-2) 

Thus, the frequency function /(xj is proportional to the expected frequency 
of the event x *= x t in a fixed number of observations. 

Since the values x* are distinct, the events x = X\ and x ~ x 2 are 
mutually exclusive. Hence, by total probability, the probability of 
x « x\ or x * X 2 is 


/(: ri) + /(* a). 



628 PROBABILITY [CHAP. 8 

In just the same way the probability of x *» or x 2 , or X* is 

£/(*,). (M) 

*-l 

It is often desirable to consider the probability that x will not exceed 
a given value. If x\, x 2> . . Xk are the values of which do not exceed 
t y then the probability that x < t is given by the sum (5-3). That is, 
the event u x < t” is equivalent to the event “x «* Xi, or x = x 2y . . or 
x * 2 *.” It is customary to write 

no = E /(*> (w) 

xiUkt 

for summation over the values of x t which do not exceed /. The function 
F(t) thus obtained is called the distribution function; it gives the probability 
that x < t. When t is so small that no x t satisfies x % < t, the sum (5-4) 
has no terms, and F(t) = 0 for such t. When t is so large that every x t 
satisfies < t , then the sum (5-4) includes every x x . In this case (54) 
gives the value F(t) * 1 . 

For example, if x is the number of heads obtained when 4 coins are tossed, 
the distribution function is described by the following table: 


t 

t < 0 

0 < t < 1 

1 < t < 2 

2 < t < 3 

3 < t < 4 

4 < t 

F(t) 

0 

He 1 

He 

‘He 

‘•He 

1 


These entries are obtained by adding the values of f(x) which were found 
previously. For instance, 1 }{q corresponds to the interval 2 < t < 3 
because 

£ f(*i) - m + fW + /(2) - M's + Me + Me - 3 Me- 

XI << 

The value A Me is the probability of getting at most 2 heads when 4 coins 
are tossed. 

The variables x considered so far in this chapter are called discrete 
variables because they assume isolated values only. For instance, the 
number of heads obtained when several coins are tossed is an integer 0, 
1, 2, 3, ... (and cannot fill up an interval). The distribution of such a 
variable is called a discrete distribution; it is defined for all values of x, 
not only for the discrete set of possible values x k . One may also think of 
the frequency function as being defined for all z, taking fix) — 0 for 
values x other than the x k , (For example, the probability of getting 3.2 
heads is zero.) The fact is that we may define fix) in any arbitrary fashion 
for values other than the x k) provided some care is taken in the interpreta- 
tion of the results. This possibility is exploited in the following discussion. 



FUNDAMENTALS OF PROBABILITY THEORY 


629 


SEC. 6) 

Graphical representation of the functions fix) and F(x) is given in 
Pigs. 4 to 6. Figure 4 is valid as a probability for all x. The relationship 
of fix) and F(x) is clarified, how- 
ever, if f(x) is modified as shown in 
Fig. 5. Here, the value of f(x) at 
any integer m is used for f(x) in the 
interval of length 1 centered about 
m. The resulting step function still 
gives the probability that x = Xk, 
provided Xk is an integer. The ad- 
vantage of redefining fix) in this 
fashion rests upon the following 
property, which is easily verified: If t is an integer, then Fit) is the area 
under the curve of Fig . 5 up to the value x « t + x /i. For instance, the 
area up to the value x ~ 2 l /i is found to be 

m +fi 1) +/(2) - F(. 2) 





Fig. 6 


630 


PROBABILITY 


[CHAP, 8 

by adding the areas of the shaded rectangles. When the values Zk are 
equally spaced, similar considerations apply to any distribution and 
frequency functions F and/. For unit spacing, that is, for Xk+i — %k * 1, 
Eq. (5-1) expresses the fact that the area under the curve is 1. 

Actually, it is possible to describe the relationship of / and F directly, without intro- 
duction of the intermediate curve (Fig. 5) . The description involves the so-called Stieltjes 
integral, which is now to be defined. Let F(t) be a nondecreasing function on an interval 
ft < t <J &» and let 4>(t) be continuous. Choose a set of points fa, t a , . . tn on the interval 
and choose intermediate values 

fa ^ £/r ^ fa- j-1* 

As the subdivision given by the fas is made finer and finer, in such a way that 

max lfa +1 - fa I -► 0, 
it can be shown that the expression 

- F(fa) 1 

tends to a limit (independent of the manner of subdivision and of the points £*)• The 
limit is called the Stieltjes integral of <t> with respect to F and is written 

\\{t) dF(t). 

Ja 

When F{t) is a discrete distribution corresponding to Xk and f(x), the function F(t) 
has a jump of value /(j*) at each value Xk but is constant between those values. Hence 
the differences 

F(ti+ i) — F(tk) 

behave much like the function exemplified in Fig. 4. They assume the value f(xk) if the 
interval (fa,fa+i) contains a single point x and they assume the value 0 if the interval 
contains no point xu- The relationship of / and F is now described by the equation 

m - f dh\x) 

where the integral is a Stieltjes integral. Although we have not defined a differential 
dF, we may think of dF(x) as being equivalent to the frequency function f{x) in the 
sense described above. 

Example 1. In terms of the distribution function, express the probability that 
a < x <b, where a and b are two numbers with a < b. 

The event “x < b” can materialize in the mutually exclusive forms 

x < a or a < x < b. 

Hence, by total probability, 

Pr (x < b) « Pr (x < a) -f Pr (a < x < b) 
where Pr means “the probability that.” This yields the desired expression 
Pr (a < x < 6) « F(b) - F(a) 
when we recall that the distribution function F(t) satisfies 

Pr {x <, 6) « F(b\ Pr (x £ a) m F( a ). 


(W) 



SEC. 6 ] FUNDAMENTALS OF PROBABILITY THEORY 631 

Example 2. In terms of the frequency function /(x) the expectation of any variable 
V - is 

E{y) « 'Zykjixk), Vk « g(xk)> (5-6) 


We consider the variable y to be defined on a sample space whose points are the n 
events 


X » Ji, X ** X2, . . , , X ** X». 


The probability of the event “x * x*” is /(x*); the value of y corresponding to the event 
“z * x*” is — 0(x*). Hence, (5-6) follows from the general definition of expectation. 


PROBLEMS 

1. Suppose a coin is tossed 5 times. What is the probability that this experiment 
will yield 0, 1, 2, 3, 4, 5 heads? 

2. If x is the number of heads in Prob. 1, make a table representing the frequency 
function /(x). Plot f(x) and also the step-function modification (see Figs. 4 and 5) 

3 . In Prob. 1 make a table and also a graph for the distiibution function F((), 

4. If a coin is tossed 5 times, find the probability that the number of heads x satis- 
fies 1 < x < 4 by use of (a) the frequency function /(r) computed in Prob. 2, (b) the 
distribution function F(t) computed iri Prob 3, (r) the stop-function graph obtained m 
Prob. 2, with reference to an appropriate area under the curve. 

6. Continuous Distributions. Since measurements are made only to a 
certain number of significant figures, the variables which arise as the result 
of an experiment are discrete. For example, if the diameter of a shaft is 
measured to the nearest 0.01 in., the measurement is a variable which 
assumes only isolated values, such as 3.21, 3 22, 3.23, ... in. Nevertheless 
it is convenient to introduce continuous variables, because they are easier 1 
to handle analytically. Such variables are now to be discussed. 

Let a point be chosen at random on the interval 0 < x < 3. How shall 
we measure the probabilities associated with that event? If the interval 
(0,1) is divided into a number of subintervals, each of length Ax = 0.1, then 
the point x is equally likely to be in any of these subintervals (Fig. 7). 
The probability that 2 0.5 < x < 0.8, 

, . . _ _ . 1 I ♦— I ) ’« I 

for example, is 0.3, since there are o 0.5 l 

three favorable cases. The probabil- p ia 7 

ity that 0.52 < x < 0.84 is found to 

be 0.84 — 0.52 = 0.32 when we divide the interval into 100 parts, and so 
on. This reasoning shows that the probability for x to be in a given sub- 
interval of (0,1) is the length of that subinterval. If Pr stands for “the 

1 This remark does not justify the use of continuous variables in applied mathematics. 
The justification rests upon the fact that discrete variables can be approximated by 
continuous ones within the experimental error. 

* In this section it will not matter w hether the intervals include their end points or 
not. Thus, Pr (a < x £ b) » Pr (a < x < b). 



632 PROBABILITY [CHAP. S 

probability that,” then 

Pr (o < x < b) - b — a, 0 < a < b < 1. (6-1) 

When (6-1) holds, the variable x is said to be uniformly distributed on 
the interval 0 < x < 1. Since the expression (0-1) may be written 

Pr (o < a- < 6) = f * dx = /* 1 dx, (6-2) 

Ja Ja 

it is customary to speak of the probability density , which in this case is 
unit} . 

More generally, a variable may be distributed with an arbitrary density 
f(x). For such a variable the expression 

m & 

measures, approximately, the probability that a: is on the interval 

t < x < t + At. 

An exact expression for the probability that £ is on a given interval (a, b) 
is 1 

Pr (a < x < 6) ~ [ b f(x)dx. (6-3) 

Ja 

This relation is illustrated in Fig. 8. 

As indicated above, the function 
f(x) is called the probability density; 
the function 

m « f /(*> dx (6-4) 

is called the distribution function. 
Evidently, F(t) is the probability 
that x is in the interval ( — »,/); 
in other words, 

F(t) = Pr (x<t). (6-5) 

Fig. 8 If/M is continuous, then (6-4) gives 

F'(t) « f(t) 

and one may speak of a probability differential 

dF(t) *= f(t) dt 

} The symbol x in (6-3) is used in two different senses. On the left x is a random varia- 
ble, and on the right x is the variable of integration. The integral could have been 

written f f(Q d£, for example. 




FUNDAMENTALS OF PROBABILITY THEORY 


SEO. 0] 

To find the distribution function t associated with the uniform density /(a?) on the 
interval (0,1) we take 

/(*) « 0, x<0 f 
f(x) - 1 , 0 < » < 1 , 

f(x) - 0, x > 1. 

This expresses the fact that x is sure to be in the interval (0,1), and is uniformly dis- 
tiibuted on that interval Hence, for 0 < t < 1, 

m - J‘ nx) dx 

- f f{x) dx +J'f(x)dx 


In a similar manner one obtains 


~0 + 



t. 


(6-6) 


Fit) * 0, i < 0, 

Fit) - 1, t > 1, 

which expresses the fact that x is never <0 but is always <1. 


The following density functions arise in many applications. 
Poisson : 


e m* , 0 < x < <», n > 0, r — nonnegative integer, 

r! 

Gauss: 

1 -h(z=*Y 

— — € Vtr/ , — OO <J<OC jC r>0, — 00<^<QO ? 

V 2tt <r 


Maxwell-Boltzmann : 


4 a 



0 < x < oo, a > 0. 


The random variable is x; the parameters r, a, <r are constants. For 
example, in the Maxwell-Boltzmann distribution x is the magnitude of 
the velocity of a gas molecule and a — m/2kT, where m is the mass, T 
is the temperature, and k is called the Boltzmann constant. A graph of 
the function for a » 1 is given in Fig. 8. The Poisson distribution is 
discussed in Sec. 11; the Gauss distribution in Secs. 9, 10, and 12. The 
latter is often called the normal distribution, but in this text the term 
normal distribution is applied to the case o * 1, u « 0 only. 



PROBABILITY 


634 


[chap. 8 


Densities and distribution functions are easily defined for several vari- 
ables. We say that f{x,y) is the probability density (or the joint proba- 
bility density) for (x,y) if the probability that (x,y) is in any given region 
R of the xy plane is 

Pr l(x,y) in R] = jj R f(x,y) dx dy. (6-7) 

The distribution function is 


F(s,t) = / / f(x,y) dx dy 

J — 00 J — GC 


« Pr [x < 8 and y < 


(6-8) 


Since probabilities are nonnegative, the density functions in (6-3) and 
(6-7) satisfy 

fix) > 0, f{x >V ) > 0. (6-9) 

Since the variables always have some finite value, in (6-3) 


and in (6-7) 


f(x) dx y 

-00 

n oo 

f(x,y) dx dy = 1. 

„ ~00 


( 6 - 10 ) 

(6-11) 


Any integrable function f(x) or fix.y) which satisfies these conditions 
(6-9) to (6-11) may be regarded as a probability density. The sample 
space is infinite; it consists of the events x = x 0 for every choice of x 0 
or (z,y) = (,x 0 ,yo) for every choice of (x 0 ,y 0 ). 

For example, if f(x,y) * l/A in a region R of area A and f(x,y) -* 0 elsewhere, it is 
easily verified that (6-11) holds. The probability that ( z,y ) is in a subregion R\ con- 
tained in R is 

Jf R fi*>y) dx d y * Jf R \ dx d v * 

where Ai is the area of R%. The variable (x,y) is then said to be uniformly distributed in R. 

The theory for finite sample spaces applies with little change to con- 
tinuous distributions; for example, the expectation is defined by 

/ <*> 

xfix) dx 

E{x) « / rf{x) dx = 

" / fix) dx 

The latter expression follows from (6-10); it shows that E(x) is the x 
coordinate of the center of mass for the area bounded by the curve y - f{x) 



SEC. 6 ] FUNDAMENTALS OF PttOB ABILITY THEORY 635 

and the x axis. More generally, the expected value of any function y « g{x) 
is 

f Vf(x) dx, y » g(x), ( 6 - 12 ) 

and the sum theorem E(x + y) ~ E(x) + E(y) is a simple consequence 
of the properties of integrals. Compare Sec. 5, Example 2. 

Two variables x, y are said to be independent if the joint density f(x,y) 
has the form 

/Cr,y) = f(x)g(y). 

The theorem of compound probability for independent events 
valid in the form 

Pr (a < x < b, e < y < d) = Pr (a < x < b) Pr (c < y < d). 

The theorem of total probability assumes various foitns, such as 
Pr (a < x < c) ~ Pr (a < x < b) + Pr (b < x < c) 
for a < b < c. Equation (6-15) is equivalent to 

/ f(x) dx — f /(;r) dx + f f(x) dx 

Ja Ja Jb 

which, in turn, is a known property of integrals. 1 

Example 1. A variable 1 x is said to be uniformly distributed on (a, b) if f(x) is constant 
on (a, b) and zero outside (a y b). Find f(x) in this case. 

Denoting the constant by c , we have 

f(x) dx *■ / c dx — c(b — a) « 1 

i Ja 

by (6-10). Solving for c yields 

/M * : » a <x <b t 

b — a 

f(x) * 0, elsewhere. 

Example 2. A stick of length a is broken at random into two pieces. Find the dis- 
tribution function F(s) for the length s of 
the shorter piece. From this find the 
probability density /(/) for the length l of 
the longer piece. 

Evidently 0 < s < a/2 in every case 
For any t between 0 and a/2 we have 
s < t if, and only if, x is on one of the 
intervals (0,0 or (a — t , a) (see Fig. 9). 
uniformly distributed, and hence 

1 It is also possible to start with the theorem of total probability and deduce from this 
that probability can be represented as a Stieltjes integral (Sec. 5). Mild continuity 
conditions then give the representation (6-3). 


mmm- 


h — * — 

V///////A 


Fig. 9 

The probability of that is 2 t/a i since x is 



( 6 - 13 ) 
is then 

( 6 - 14 ) 

( 6 - 15 ) 



PROBABILITY 


[CHAP. 8 


m - 

a 

°<*^- 

m - o. 

» < 0, 

m - 1, 

a 

8> 2' 

Since the length l of the longer piece satisfies l » a - s, we have l < t if, and only if, 
« > a — t. By the result just obtained the probability is 

1 — Pr (s < a — 

0-1 2(0-0 
a 

for a/2 < t < a and 0 or 1 otherwise. This gives the distribution function for L By 
differentiation, the density is found to be 

*-!• 

a 

2 < 1 < 

m - o, 

(6-16) 

elsewhere. 


The differentiation is not valid for l ~ a/2 or for l ** a, but it does not matter how the 
density is defined at these isolated points. 

Example 3. A stick of length a is broken at random, and the longer piece is again 
broken. What is the probability that the three segments can form a triangle? 

Let l be the length of the longer piece. If this piece is broken at a point x , the three 
segments are a — l, x, l — x. The condition for a triangle is that the sum of any two 
segments shall exceed the third: 

a — l + x > l — x t a — x > x, l > a — l. 

Since l > a/2 automatically, these conditions reduce to 

( 6 * 17 ) 

It is a conceptual aid (and not incorrect) to use the theorems of total and compound 
probability in the following manner: The probability that l is on the interval (l, l + dl) is 



a 


by (6-16). 


since x is 
product 


After l is chosen, the probability that x satisfies (6-17) is 

f ' 2 i £ 

Jl-al 2 l 


0/2 l dx - — * 


uniformly distributed on (0,f). The probability of both these events is the 
a -12 


l a 


dl 


by compound probability, and total probability now gives the final answer: 



SEC. 7] PROBABILITY AND RELATIVE FREQUENCY 637 

Example 4. Buff on's Needle Problem, A needle of length a is dropped on a board 


which is covered with parallel lines spaced 
probability that the needle intersects one 
of the lines? 

We assume that the variables x and $ 
of the figure are uniformly distributed, 
s being the distance from the center to 
the nearest line. There is intersection if, 
and only if, |(o/2) cos 0| > x. For fixed 6, 
the probability of this is 

| (a/2) cos 6 1 o 1 cos 6 1 
6/2 ~ b 

since x is uniformly distributed on (0,6/2). 
Using total and compound probability as 
in Example 3, we obtain the final answer: 

2 * ajcoagj d$ a 1 
b 2t " 6 2ir 


a distance 6 > a (Fig. 10). What is the 



r*' 2 2a 

4 / cos 9 dd m . 
Jo TO 


PROBLEMS 


1. A probability density is defined by f(x) « 3x 2 for 0 < x < 1 and f(x) « 0 else- 
where. Find E(x) and #(x 2 ). Find the distribution function F(x), and from this obtain 
a value m such that x is just as likely as not to exceed m. (The value tn is called the 
median of x.) 

2 . The radius of a sphere is uniformly distributed on (0,1). Find the expected value 
of the volume (see (6-12)]. What is the probability that the volume exceeds half its 
maximum value? 

3 . A stick of length a is broken at random into two parts. What is the expected 
length of the shorter part? 

4 . Two points are chosen at random on a line of length a . What is the probability 
that the three segments can form a triangle? 

6. The probability density for bullets hitting a target is given by 




2t Cgffp 


where <r x , <r V) m Xt my are constant. Sketch the curves of constant density in the xy plane. 
What kind of curves are they? 

6. We make two independent observations xi, x 2 of a variable with distribution func- 
tion }{x). What is the probability that a third independent observation j 3 will fall 
between x\ and £ 2 ? Generalise to n observations. Hint; Use the methods of discrete 
probability. 


PROBABILITY AND RELATIVE FREQUENCY 

7. Independent Trials. It often happens that the probability of an event 
cannot be determined by counting cases or by other a priori considerations. 
Sometimes the determination is impossible in principle; for instance, 



638 PROBABILITY [CHAP. 8 

one cannot compute the probabilities associated with a loaded die or the 
probability that a given radio tube will fail in the first hundred hours’ 
use. Sometimes the determination is theoretically possible but impractical 
For instance, by examining every nail in a 100-lb keg one could find the 
probability that a nail selected at random will be defective, but this is 
not a useful method. 

In many such cases an estimate for the probability can be obtained by 
repeated trials (or by inspecting a suitable sample , in the terminology of 
statistics). In the case of a biased coin, for example, if 10 tosses give 7 
heads, 100 tosses give 73 heads, and 1 ,000 tosses give 090 heads, it appears 
that “the probability of heads is probably close to 0.7/’ The two italicized 
words express a reservation which is always present in conclusions such 
as this. 

The figures 7, 73, 690 in the above discussion represent the frequency 
of heads; the ratios 

7/10, 73/100, 690/1,000 

give the relative frequency in 10, 100, or 1,000 trials. More generally, 
if an event occurs m times in n trials , the relative frequency is m/n , 

The trials in a sequence of trials are said to be independent if the proba- 
bilities associated with a given trial do not depend on the results of pre- 
ceding trials. For example, the probability of heads on a given toss of a 
symmetric coin is no matter what is known about the results of previous 
tosses. But if we try to get an ace by drawing cards one at a time without 
replacing, the trials are dependent. In this case, the probability of ace 
in a given trial depends on the number of aces that may have been drawn 
previously. 

When an event has constant probability p of success, the probability 
of m successes in n independent trials may be computed as follows. A 
sequence of m successes and n — m failures is represented by a sequence 
of m letters S and n — m letters F: 

SSFFSS . . . SF. (7-1) 

Since the trials are independent, the probability of any one such sequence 
is 

ppqqpp . . . pq « p m q n ^ m } (7-2) 

where q ** 1 — p. To obtain the number of favorable sequences, observe 
that a sequence is determined as soon as the positions of the m letters S 
are fixed. The m places for these letters S can be chosen from the n places 
in n C m ways, and hence the required probability is 

ptyi-m + p ™ q n~™ + . . . + 

by the theorem of total probability. 


( 7 - 3 ) 



PROBABILITY AND RELATIVE FREQUENCY 


639 


sec. 7] 

Alternatively, the reader may imagine a sample space in which each event consists of 
a sequence (7-1), with associated measure (7-2). Then (7-3) represents the sum of the 
measures of those points favorable to the event: m successes. 

Replacing m by x gives 

B(x) - n C x p*q n ~* m — — — p*(l - p) n ~* (7-4) 

x!(n ~ x ) ! 

for the probability of exactly x successes in n independent trials with 
constant probability p . The associated distribution function is 

F(t) « n Coq n + nOm"- 1 + • • • + n(Vg w ~‘ (7-5) 


for integral values of t. This expression gives the probability of getting 
at most t successes in n trials. 

Because of its connection with the binomial theorem (Prob. 6), the 
function B(x) is called the binomial frequency function, F(t) in (7-5) is the 
binomial distribution , and the statement that B(x) gives the probability 
of x successes in n independent trials is called the binomial law of proba- 
bility. Since many statistical studies involve repeated trials, the binomial 
law has great practical importance. 

To illustrate the use of the formula (7-4) let it be required to find the 
probability that the ace will appear exactly 4 times in the course of 10 
throws of a die. Here p = Q ~ n — 10, x - 4. Hence the proba- 


bility is 


B( 4) 


io! ziy/sy 

4!G!\G/ W 


0.05427. 


Since the expected number of successes in one trial is p , the expected 

number in n trials is N ^ . 

L(x) - np (7-6) 

(compare (5-2)]. For most distributions there is no special relation between 
the expected value and the most probable value, but for the binomial dis- 
tribution they happen to be almost equal. Equation (7-4) yields 

B(x + 1) (n — x)p 

B(x) (x + 1 )q 

after slight simplification. Hence B(x) is an increasing function of the 
integer x if, and only if, 

(n — x)p 

- — > 1 . 

(x + 1 )q 

The latter inequality is the same as 

(n — x)p > (x + l)g 

which reduces to np > x + q, since p + q * l. We have shown, then, 
that B(x + 1) > B(x) as long as x < np — q but B(x + 1) < B(x) there- 



640 


PROBABILITY 


(CHAP. 8 


After. Since q < 1, this establishes that B(x) is maximum for a value of 
x which is within 1 of the value x np. Further discussion of the func- 
tion B(x) is given in the following sections. 

Example 1 . Ten tosses of a suspected die gave the result 1 , 1 , 1 , 6 , 1 # 1 , 3, 1, 1, 4. What 
is the probability of at least this many aces if the die is true? 

The event “at least 7 aces” can materialize in four mutually exclusive ways: 7 aces, 
8 aces, 9 aces, 10 aces. By total probability (or by use of the distribution function) the 
required answer is found to be 


B( 7) + B(8) 4* B( 9) + B(10) 

~ i oC 7 (K ) 7 (%) 3 + xo C 8 (K ) 8 (^) 2 + + xo C 10 (M ) 10 

when we take p » K, n « 10. This reduces to 0.00027, approximately. Because the 
observed result has such small probability, one would reject the hypothesis tl p *» 
unless there is some other evidence in its favor. 

Example 2. In Example 1 let p be the unknown probability of the ace in a toss of 
the die. (a) For what value of p does the expected number of aces agree with the ob- 
served number? ( b ) For what value of p is the probability of the observed result a maxi- 
mum? 

Since E(x) -* np by (7-6), the observed and expected numbers agree when p *•» x/n, 
that is, when p » 0.7. The estimate for p given by p — x/n is called an unbiased esti- 
mate, because E{x/n) « p. 

For part (b), the probability of getting 7 aces and 3 other numbers is 
p 1 q 8 or igCtpV, 0-1 ~ P, 


depending upon whether the order is considered or not. In either case the probability 
is maximum when p 7 (l — p) 3 is maximum. This, in turn, is maximum when 

log p 7 (l - p) 3 » 7 log p + 3 log (1 - p) 

is maximum. Differentiation gives 

- _ _JL_ _ o, 

p I - p 


or p *» 0.7. An estimate for p such as this, which maximizes the probability of the ob- 
served result, is called a maximum likelihood estimate. 


PROBLEMS 

1. When 5 coins are tossed what is the probability of exactly 2 heads? At least 2 
heads? What is the expected number of heads? The most probable number of heads? 

2. If 6 dice are tossed simultaneously, what is the probability that (a) exactly 3 of 
them turn the ace up? {b) At least 3 turn the ace up? 

3 . If the probability that a man aged sixty will live to be seventy is 0.65, what is the 
probability that out of 10 men now sixty at least 7 will live to be seventy? 

4 . A man is promised $1 for each ace in excess of 1 that appears in 6 consecutive 
throws of a die. What is the value of his expectation? 

5. A bag contains 20 black balls and 15 white balls. What is the chance that at 
least 4 in a sample of 5 balls are black? 

0 . (a) By use of the binomial theorem show that 



BBC. 8] 


641 


PROBABILITY AND RELATIVE FREQUENCY 

[q «f pf) n - B( 0) + £(1)* + B(2)fi + * - + B{n)t\ 

( b ) Interpret the h entity which arises when t « 1. (c) Differentiate with respect to f, 
and interpret the identity which then arises for t «• 1. [The function (g 4- p0 n is 
called the generating function of the sequence j£(x) } .] 

7. (a) One hundred light bulbs were tested for 500 hr, at the end of which time 57 
bulbs had failed. Obtain an unbiased estimate and also a maximum-likelihood esti- 
mate for the probability of failure in 500 hr. (6) Are these two estimates of p always 
equal for the binomial distribution? Hint: In (5), compare the result of maximizing 
p m q n ~ m with respect to p and the result of choosing p so that E(x) ■» m, where m is the 
number of observed successes. 

8. In a certain agricultural experiment, the probability that a plant will have yellow 
flowers is If 10,000 plants are grown, what is the probability that the number with 
yellow flowers will be between 7,400 and 7,000? (To appreciate later developments ob- 
serve that your answer, which should be indicated only, is difficult to compute.) 

8. An Illustration. Some interesting conclusions concerning the bino- 
mial law are suggested by an example that presents many features of the 
general case. Consider a purse in which are placed 2 silver and 3 gold 
coins, and let it be required to find the probability of drawing exactly x 
silver coins in n trials, the coin being replaced after each drawing. The 
probability of exactly x successes in n trials is given by (7-4) where p, 
the probability of drawing a silver coin in a single trial, is %. If the 
number of drawings is taken as n — 5, 10, or 30, the respective frequency 
functions B(x) are 


B(x) - 6 C x (H) x (H) s ~ x , 

3 

11 

Ol 

B(t) = ,0 c x (%) x (K) 10 -*, 

n = 10, 

B{x) = 3 oC x (H) z (H) M ~ x , 

n — 30. 


By use of these expressions one can compute the values of B(x) to any desired accu- 
racy. The result of such a computation to four places of decimals is presented in the 
accompanying tables. In the third table the entry 0.0000 is made for 0 < x < 2 and 
for x >23 because in these cases B(x) was found to be less than 0.00005. For example, 
the probability of drawing exactly 23 silver coins in 30 trials is 

£(23) - «C„($$)“(?$) 7 - 0.000040128. 

The reader can verify that the most probable values of z are exactly equal to np 
(and not merely within 1 of np). This behavior is always found when np is an integer. 

Probability of Exactly x Successes in 5 Trials 


X 

B(x) 

X 

B(x) 

0 

0.0778 

3 

0.2304 

1 

0.2592 

4 

0.0768 

2 

o.$m 

5 

0.0102 



642 


PKOB ABILITY 


[CHAP. 8 


Pbobabiutt of Exactly x Successes in 10 Trials 


X 

B(x) 

X 

B(t) 

X 

B(x) 

0 

0.0060 

4 

0.2508 

8 

0.0106 

1 

0.0403 

5 

0.2007 

9 

0.0016 

2 

0.1209 

6 

0 1115 

10 

0.0001 

3 

0.2150 

7 

0.0425 

: 




Probability of Exactly x Successes in 30 Trials 


X i 

B(x) 

X 

B{x) 

X j 

Hix) 

<2 

0.0000 

9 

0.0823 

16 

0.0489 

3 

0.0003 

10 

0.1152 

17 

0 0269 

4 

0.0012 

11 

0.1396 

18 

| 0 0129 

5 

0.0041 

12 

OA474 

19 

0.0054 

6 

0.0115 

13 

0.1360 

20 

0.0020 

7 

0.0263 

14 

0.1100 

21 

0.0006 

8 

I 0 0505 

15 

0.0783 

22 

0.0002 





>23 

0.0000 


The values given in the tables are presented graphically in Fig. 11 after 
the manner described in Sec. 5. Each curve has the general shape pre- 
dicted by the theory of the preceding section, but the figure shows also 
how the shape changes as we proceed from one curve to another. The 
numerical area under each curve is 1 , although the curves become broader 
and flatter as n increases. In particular the maximum (that is, the proba- 
bility of the most probable value) decreases as n increases. This is just 
what one would expect intuitively. (For instance, one could easily get 
2 heads in 4 tosses of a coin, but one would be surprised to get exactly 
500,001 heads in 1,000,002 tosses.) The fact that the curves become 
broader indicates that the values of x experience a wider spread when 
there are more trials, and this, too, one would expect. Naturally, the 
curves ought to get broader if the maximum is to decrease while the area 
remains equal to L 

The foregoing discussion is concerned with the frequency of success in 
n trials. The results are very different if, instead, one considers the relative 
ffequency x/n. The distribution for the variable x/n is presented graphi- 
cally in Fig. 12. These curves were obtained from the preceding by the 
change of scale indicated on the axes, and hence, the area is still 1. Instead 
of becoming broader, these curves become narrower as n increases. The 



PROBABILITY AND RELATIVE FREQUENCY 


643 


SEC. 8] 

relative frequency x/n tends to cluster about its expected value p as n 
gets large. It is for this reason that relative frequency can be used to 
estimate an unknown probability. 



Fig 11 

The behavior suggested by this example may be summarized as follows. 
When the number of trials n becomes large, the absolute deviation from 
the expected value 

\x — np\ = | x — E(x) | 

inB(x) 



Fig. 12 

is likely also to be large, but the relative deviation 


x — np 

ss 

X 

p 

XK 

X ( x\ 

- - E -) 

n 


n 


n viz 


is likely to be sipalh 1 

1 It will be seen in Sec. 9 that the first expression is usually of the order \/n and the 
second, of order \/y/n\ compare Prob. 3. 



644 


PROBABILITY 


[CHAP. 8 


PROBLEMS 

1. Plot a distribution curve like that of Fig 1 1 for the probability of x successes in 4 
trials when p » %, Shade the area corresponding to the event 1 < x < 3, and find 
the probability of this event. 

2 . For plot the probability of the most probable number of successes versus to. 

(Take points at w « 1, 2, 3, 4, 5410, 30 only, cf. Prob, 1 and accompanying tables.) 
On the same figure plo t 1 /-y /2 irnpq versus to. (It is shown in Sec. 9 that the probability 
is asymptotic to \/y/2vnpq when n is large. This expression appioachos zero as to *-> *>, 
even though we are considering the most probable value ) 

3. Using the tables and your numerical values in Prob. 2, plot y/n B(x) versus 
(x — np)/y/n for p * % and for « = 3, 4, 5, 10. Use the same scale in each case. 
Formulate a conjecture concerning the behavior as to ~+ and test your conjecture 
by plotting five well-chosen points on the curve corresponding to n ** 30. 

9. The Laplace-de Moivre Limit Theorem. Numerical computation of 
the binomial distribution is difficult when n is large. In this section an 
approximate formula is obtained when n and np are both large. In Sec. 11 
a formula is found when n is large but np is not large. These approxima- 
tions, together with the exact formula when n is moderate, cover all cases. 

The analysis is based on the Stirling formula , 

n! ~ n n e~ n \^2rn, (9-1) 

which is made plausible by the following dis- 
cussion. Consider the function y = log#, 
and observe that for k > 2, 



f 

h - 1 


log x dx> ^[log (As — 1) + log k), 


log#. 


since the right-hand member represents the 
trapezoidal area formed by the chord (Fig. 13) 
joining the points P and Q on the curve 
Denote the area between the chord and the curve by so that 


r log x dx 

-i 


HHog {k - 1) 4 log As) 4- a*. 


(9-2) 


Setting k — 2, 3, . . . , to in (9-2) and adding give 

J log x dx * H(log l -f log 2) -f H ( log 2 4 log 3) -4 

4 J 4 [log (n — 1) 4* log n] 4 (os 4 as 4 h o»). 

Integrating the left-hand member and combining the terms of the right-hand member give 

» 

n log n - n 4“ 1 « log n! - H log n 4 X) a*. 

t **2 


log n! - (n 4 H) log to - n 4 1 — X) (H . 

i-2 


Hence, 


(9*3) 



645 


SBC. 9] PROBABILITY AND RELATIVE FREQUENCY 

Since each 04 is positive, it follows that 

log n! < (n H) log n — n -f 1 

and hence 

n! < eVn^e"*, (9-4) 

The expression on the right of the inequality (9-4) is, therefore, an upper bound for n!. 
To get a lower bound, solve (9-2) for a*, perform the integration, and obtain 




Now, since the integrand is nonnegative, 


r (i-iy 

Jk - 1 Vx k/ 


dx > 0 


and the evaluation of (9-6) leads to the formula 


log- 


2k 


k — 1 2 k(k - 1) 

By use of this inequality, (9-5) gives 

1 

die < 

4 k{k - 1) 4 1 

— 5'<if("D + G-3*-+(.-iT-i)K- 


-(-± — i). 

4 \k - 1 k/ 


<m> 

(9-6) 


By means of this result and (9-3), one obtains 

log n! > (n -f Vi) log n — n + 1 — 

whence n! > e*A a / n n n e~ n . (9-7) 

Combining (9-4) and (9-7) furnishes the inequality 1 

f^Vn rt n c*" n < «* < cVn n n e~ n 


for all values of n > 1. Since e «® 2 718, e* 4 = 2 1 17, and «= 2.507, we have shown 
that (9-1) is correct as to ordei of magnitude. More refined methods establish that the 
error is less than 10 per cent for n > 1, less than l per cent for n > 10, and less than 
0.1 per cent for n > 100. Moreover, the percentage error approaches zero as n -+ oo, 
so that the equality is asymptotic. 

In the expression 

B (r) = — PY~ T (9-8) 

r!(n — r)l 


for the probability of r successes in n independent trials, we assume that 
r, n, and n — r are large enough to permit the use of Stirling’s formula 

1 The derivation of this result is given by P. M. Hummel, Am. Math. Monthly , 47:97 
(1940). 



646 


PROBABILITY 


[CHAP. 8 


(9-1). Replacing n!, r!, and (n — r)! by their approximations gives, after 
simplification, 




r) 


(9-9) 


Let $ denote the deviation of r from the expected value np; that is, 

5 = r — np . 
n — r — nq — 8 

(np-f5) / i) 

nq) 


Then, 

and (9-9) becomes 1 


B(r) 


or 


where 


^L w (i + i)(i - ±) 

i+— ) (l ) 

np/ \ nq/ 


\ np/ \ ntf/ 


Then, log £(r)A ~ — (np + S) log (l H ) — (nq — 6) log ( 1 ~ 

\ np/ \ nq/ 


Assuming 1 6 1 < npq f so that 


5 

< i 

and 


np 



nq 


< 1, 


permits one to write the two convergent series 


and 


Hence, 


log (l H \ = 

\ np/ 

log(l~-) = 

\ nq/ 


b 

np 


+ 


S 3 


2 n*p 2 3 n 6 p 


3^3 


nq, 

log B(r)A ^ — 


5 r 

n<? 2n 2 g 2 3n 3 <? 3 

JL _ g3 (p 2 ~ g 2 ) 

2npq 2-3n 2 p 2 q 2 


<V + 9 S ) 

3 • 4n 3 p 3 9 3 


Now, if 1 5 1 is so small in comparison with npq that one can neglect all 
terms in this expansion beyond the first and can replace A by \^2mpq, 

1 Here and in similar eases which arise subsequently, we assume that p & 0 and 
q s* 0. The cases p 0 or p « 1 can be dealt with by inspection. 



647 


SBC. 9 ] PROBABILITY AND RELATIVE FREQUENCY 

then there results the approximate formula 


B(r) 


— e 


— !*/2 npq 


( 9 - 10 ) 


\/ 2tt npq 

which bears the name of Laplace’s, or the normal, approximation . With 
cr = V npq , Eq. (9-10) becomes 

1 


B(r): 


■\/ 2 t a 




(9-11) 


The equality is asymptotic; that is, the ratio of the two sides tends to 
1 as n — * co. A comparison of B(r) with the normal approximation is 
given in Fig. 14. 



The main usefulness of this result is to compute the probability 

E *00 0-12) 

r«r i 

that the number of successes is between the given limits r x and r 2 . Equa- 
tion (9-11) shows that the sum (9-12) may be approximated by a sum 

S — Lr- e-**' 2 ** (9-13) 

V2ir a 


over appropriate values of 8. Since 5 — r — np, the difference between 
successive values of 8 is 1, and hence if \\e let t — 8/ a, the difference be- 
tween successive values of i is At = 1/ cr. Thus (9-13) becomes a sum over t, 


\/2r 


-t*f 2 


AL 


(9-14) 



PROBABILITY 


648 PROBABILITY [CHAP. 8 

As At 0, the expression (9-14) approaches an integral, which may be 
evaluated in terms of the function 

1 


* (< > ={ 




e -t*(2 dt 


(9-15) 


tabulated in Appendix D. These considerations yield the following funda- 
mental result, known as the Laplace-de Moivre limit theorem : 

Theorem. Let x be the number of successes in n independent trials with 
constant probability p. Then the probability of the inequality 


x — np 

t\ ^ < (g 

‘ Vnpq 

approaches the limit 

~ /%-**» eft - *(t 2 ) - Hh) 

\/2ir J t i 

as n — > oo. 


(9-16) 

(9-17) 


To complete the proof one must note that the error in passing from (9-12) to (9-13) 
ifl small for large n even when the number of terms in the sum is large A more de- 
tailed analysis, taking due account of this question, is given in William Feller, “Proba- 
bility Theory and Its Applications/' pp. 1 33—137, John Wiley Sons, Inc., New York, 
1950, It is shown that a better approximation is given by 




Vnpg, 


(9-18) 


although the improvement is not important when n is large An expression for the 
error in the approximation is derived in J V. Uspensky, “Introduction to Mathematical 
Probability/' p. 129, McGraw-Hill Book Company, Inc , New York, 1937 


To illustrate the use of the result (9-17), let us find the probability that 
the number of aces will be between 80 and 110 when a true die is tossed 
600 times. Here n — 600, p - Y§, q = and x varies from 80 to 110. 
Hence 


80 - 100 

t X = — t t-tt == •■ r=r = —2.19 and 

v(100)(%) 


110 - 100 

h ~ “U(iooW) 


1.09. 


The table gives $(f 2 ) - 4>(1.09) ~ 0.362, and similarly 


4>(— 2.19) * -$(2.19) = -0.486. 

[Observe that 4>( — t) - — $(/), wince the curve y = e “** s/2 is symmetric.] 
Hence the required probability is, approximately, 

0.362 - (-0.486) - 0.848. 

Example 1. In the notation of the text, the probability P inax of the most probable 
value of r satisfies 

i P max ' N “ / 


whan n is large. 


y/2rnpq 


(9-19) 



SEC. 9] PROBABILITY AND RELATIVE FREQUENCY 649 

In Sec. 7 it was found that the most probable value is of the form 
r «* np -f 6, |0| < L 

For this value of r we have 

h » r — np » 0 

and hence Eq, (9-10) shows tlrnt the associated probability is asymptotic to 


. g— 4* i'lnpq^ 

y/2irnpq 

As n oo, the exponential tends to 1, since is bounded, and this yields (9-19). 

Example 2. In an agricultural experiment Mendelian theory yields a probability 
p * that any given plant should have blue flowers. Out of 10,000 plants it was 
found that 2,578 had blue flowers. Does this result contradict the theory? 

According to theory, there should have been 2,500 plants with blue flowers; that is, 
the expected number is np » 2,500. There were, in fact, 78 more than this. We have 
to decide if this excess is too large to be attributed to chance. 

Let us find the probability that the excess will be 78 or more if the hypothesis p m 
is indeed correct. The inequality 

78 < x - np (9-20) 


becomes 

when divided by y/ npq 


r — nv 

1.801 < < oo (9-21) 

V npq 

43.3. According to the table the probability of (9-21) is 


#(*>) - <£(1 .801) « 0 500 - 0 464 * 0.036. 


Now, in a statistical tost it is customary to reject the hypothesis if the hypothesis makes 
the probability of the observed result levs than a fixed quantity a determined before- 
hand The value a (which is called the significance level of the test) is of km taken to be 
0 05. Since our probability 0 036 is less than 0 05, the experimental outcome is con- 
sidered too unlikely to be attributed to chance, and we reject the hypothesis (( p =» 

In this sense, the experiment contradicts Mendelian theory. 

We now give another analysis which leads to the opposite conclusion. Instead of 
saying ‘The excess was 78,” one could just as w r ell say, “the discrepancy w as 78,” meaning 

\x — np| 78. (9-22) 

Both statements are equally valid descriptions of the experimental outcome. The 
probability of 

\x — ?ip| < 78 

is found, as above, to be 

#(1.801) - #(-1.801) - 0.928, 


and hence the probability of the contrary event is 

1 - 0 928 - 0.072, 


Since 0.072 > 0.05, a discrepancy of ‘78 or more” is sufficiently probable to be at- 
tributed to chance (if, as before, our significance level is 0.05). Hence the hypothesis 
is not contradicted by the experiment. 1 

1 When the probability exceeds the significance level, as in this case, the hypothesis is 
not thereby proved but it is considered to have withstood the experimental test. 



650 


PROBABILITY 


[CHAP. 8 

It requires statistical methods of considerable subtlety to decide between competing 
tests of a hypothesis such as the foregoing. These methods show that the first procedure 
is appropriate for testing the hypothesis “p — against the alternative u p > l /i" 
whereas the second is appropriate for testing the hypothesis against the alternative 
u p & A very readable account of the subject is given In P« G. Hoel, “Introduction 
to Mathematical Statistics,” chap. 10, John Wiley & Sons, Inc., New York, 1954. 


PROBLEMS 


1. Two dice are tossed 1,000 times. What is, approximately, the probability of get- 
ting a sum of 4 the most probable number of times? 500 times? (l T se a table of ex- 
ponentials.) 

2 . A true coin is to be tossed 1,600 times, and it is desired to find the probability that 

the number of heads x will satisfy 780 < x < 830. (a) Show that this inequality is 

equivalent to 


-1 


< 


x — np 
y/n'pq 


< 1.5. 


(6) Express the probability of the latter inequality m terms of 4> by means of the normal 
law. (c) Using the table, evaluate the probability. 

3 . By means of the normal law, obtain an approximate numerical answer to Prob. 8, 
Sec. 7. 

4. A machine has a probability p «= 0 01 of producing a defective bottle In oik* 
day's run, out of 10,000 bottles, 120 were defective, h ind the approximate probability 
of at least this many defectives if the machine is running as usual 

5. A suspected die gave only 960 aces m 6,000 tosses. If the die is true, (a) w hat is 
the probability of getting at most 960 aces in 6,000 tosses? (b) What is the probability 
of getting a discrepancy \x — np\ of “40 or more”? (c) At a significance level of 0.05, 
does either calculation indicate that the die is loaded? 


10. The Law of Large Numbers. Since 2 B(r) = 1 for each value of n, 
it is natural to expect, by the foregoing analysis, that 


1 

V2? 



1. 


(10-1) 


For a direct proof of (10-1), define I by 


I = r e-^dx = f e^'Uy. 

J — 00 J —00 


( 10 - 2 ) 


Then multiplication of the two expressions (10-2) yields 

J 2 = f f X e-^^'Uxdy (10-3) 

J — ao J — oo 


after changing to a double integral. In polar coordinates, 


n oo /*qo 

e~ T ‘ l2 r dr d6 = 2w e ~ r ‘ 12 d (r 2 / 2) =* 2r (10-4) 



SEC, 10] PROBABILITY AND RELATIVE FREQUENCY 651 

so that I = -\/2tr, and (10-1) follows. The transformations leading to 
(10-3) and (10-4) are justified by the fact that (10-3) is an absolutely con- 
vergent double integral. 

Equation (10-1) shows that the function 


1 

-\/2t 


r e~ x * /2 dx « *(*) + - 
*' — 00 9 


is a distribution function; it is called the normal distribution . The theorem 
of the preceding section asserts that the variable h/cr is approximately 
normally distributed when n is large. This fact will now be used to es- 
tablish the following fundamental result, which is a special case of the 
so-called law of large t lumbers : 

Theorem. Let x be the number of successes in n independent trials with 
constant probability p . If e is any positive number y then the probability of the 
inequality 

I x I 


- - p 

n 


< € 


(10-5) 


tends to 1 as n — » oo. 

In other words, the relative frequency of the event is almost sure to be 
close to the probability of the event when the number of trials is large. 
For proof, write the inequality (10-5) in the form 


which becomes 


x — np 

< < e 

n 



( 10 - 6 ) 


when multiplied by \A/ pq Given any number t {) (no matter how large), 
we can choose v so that t y/n’pq > f 0 . In this case the probability of the 
inequality (10-6) is at least equal to the probability of 


As n 


x — np 

— *o < 7~ — ~ ^ ^o* 

V npq 

oo, the latter probability 1 tends to 


1 

y/Vitr 



(10-7) 


( 10 - 8 ) 


by (9-17). Since to is as large as we please, Eq. (10-1) shows that the in- 
tegral (10-8) is as close to 1 as we please, and this completes the proof. 

1 One must not apply (9-17) directly to (10-6), because (9-17) was obtained only for 
fixed t\ and h whereas the limits in (10-0) depend on n. 



602 


PROBABILITY 


(chap. 8 

The theorem was established first by James Bernoulli (1654-1705) 
after 20 years of effort. The law of large numbers lies at the basis erf all 
attempts to estimate a probability experimentally, and it affords a phil- 
osophical justification for such attempts. In fact, some developments of 
the subject define probability in terms of relative frequency, by the formula 
p = lim (x/n) as n oo, and rely on the law of large numbers to ensure 
that the limit exists. 

The theorem makes possible some interesting computational procedures, 
known as Monte Carlo methods. Although the method is not to be discussed 
at length here, we sketch an example that illustrates some of the main 
features. Suppose a man walks in a straight line, taking a step of length 
h ft every $ sec (see Fig 15). Each step is equally likely to be to the right 

h 

Fig. 15 

or to the left, without regard to the preceding steps. Assuming that x 
is a multiple of h and t is a multiple of s, it is required to find the probability 
that the man is x ft from his starting point at time t. 

Let U(x 7 t) stand for the probability in question; that is, U(x,t) is the 
probability of the man's being at point x at time t if he was at point x = 0 
at time t — 0. Now, he can arrive at point x at time t + s in two ways. 
Either he was at point x + h at time f and took a step to the left, or he 
was at point x — h at time t and took a step to the right. The probability 
of being at x + h at time t is U(x + h, t) by the definition of £/, and the 
probability of a step to the left is Jd? by hypothesis. Hence the probability 
of both events is 

y 2 u(x + h 9 1) 

by compound probability. In just the same way the probability of being 
at x — h and then stepping to the right is 

y 2 u(x - h, t). 

By total probability, the probability of getting to the point x at time 
t 4* s is the sum, and we are thus led to a difference equation for U, 

U(x, t + s) = yu(x + h,t) + yu(x - h f t). (10-9) 
The boundary conditions are 

U(x t 0) = 0 for x 0, X) U(x f t) « 1 (10-9a) 

X 

which express the fact that he is sure to be at the origin when t « 0 and 
sure to be at some point x for all t 




probability and relative frequency 


SEC* 10] 


633 


To apply the Monte Carlo method to this problem, we make a large 
number of actual random walks experimentally. The number of times 
we arrive at point x at time t gives an estimate for the probability U(x,t) 
by virtue of the law of large numbers. Hence, the calculation yields an 
approximate solution of the problem (10-9) without any direct use of 
(10-9). In practice, the “random walks” are made on a computing machine 
by reference to a set of random numbers. Similar methods apply to 
difference equations of much greater complexity than (10-9). 


For readers familiar with the theory of heat conduction the foregoing example yields 
an interesting interpretation 1 of the normal law. Subtracting U(x,t) from both Bides 
of (10-9) and dividing by s give 

U(x, t+s)~ U(x t t) h 2 * 4 U(x 4- h, t) - 2l)(x,t) 4- U(x - k , t) 

« ~2sL > J 


If we set s » h? and let h 

with boundary condition 

f/(j,0) 


0, this becomes, formally,* 

— 1 a2?/ 
dt “ 2 ax 2 


o 


for 


f U(x,t) 
J QO 


dx 


I. 


(10-10) 

(10-lOa) 


Since these are the conditions for an instantaneous source of heat at the origin, a solution 
is * 


V(x } t) 


1 




(10-11) 


Now, in the random walk the probability of a steps to the right and b to the left is 
given approximately by the normal approximation (9-10); it turns out to be 


4 


ir(a + b ) 


( a _ fe )2/2(«+fe) 


( 10 - 12 ) 


If the man arrives at point x at time t, he makes t/s steps altogether and x/h more steps 
to the right than to the left: 


a 4 ~ b 


Substitution in (10-12) and setting s ** h 2 yield 

1 




\/ 2irt 




(10-13) 


for the probability. Here 2 h is the distance between possible values of x when t is fixed, 
and hence the coefficient of 2 h may be regarded as a probability density. The condition 

1 Since heat is due to random motion of the molecules, the analogy of the random-walk 

problem with the problem of heat flow has a physical basis as well as the mathematical 

basis outlined in the text. 

4 See Chap. 6, Sec. 26. 

1 See Chap. 6, See. 19. 



654 PROBABILITY [CHAP. 8 

“h small 1 ’ means simply that the number of steps is large, so that the normal law is ap- 
plicable. The analogy between (10-11) and (10-13) is evident. 

The discussion shows that not only (10-9) but the problem in heat flow given by 
(10-10) may be attacked by making random walks. Some of the main applications of 
Monte Carlo methods are, in fact, to the study of partial differential equations. 

Example: A true coin is tossed repeatedly. It is desired to have a probability of 0.99 
that the relative frequency of heads shall be within 1 per cent of the probability of 
heads. How many times must the coin be tossed? 

If the coin is tossed n times, the desired inequality is 


which m the same as 

- 0.01 yj'- 

Setting the probability of (10-15) equal to 0 99 and noting that p * q, we get 

0.99 - *(0 0lVn) - ♦(-OOlV'n) 833 2<J>(0.01 y/n) 
by the normal approximation. The table gives 

0 01 Vn = 2.58, 

so that n 07,000 approximately. The fact that a problem such as this will always 
yield a finite value for n is the essential content of the law of large numbers. Applying 
the law of large numbers in another fashion, we ran interpret the result more or less as 
follows: If the whole coin-tossing experiment is repeated a great many times, in about 
99 per cent of these experiments the inequality (10-14) w ill be verified. 

PROBLEMS 

1. In the Example of the text, how many times must we toss the coin to make the 
probability 0.95 that the relative frequency is within 5 per cent of the probability? 

2. On the average a certain student is able to solve 60 per cent of the problems as- 
signed to him. If an examination contains 8 problems and a minimum of 5 problems 
is required for passing, what is the student’s chance of passing? Hint Because of the 
law of large numbers, you may take the statement about the student’s average per- 
formance to mean: “His probability of solving any given problem is 0,6.” 

8. If Paul hits a target 80 times out of 100 on the average and John hits it 90 times 
out of 100, what is the probability that at least one of them hits the target when they 
shoot simultaneously? 

4 . If on the average in a shipment of 10 cases of certain goods 1 case is damaged, what 
is the probability that out of 5 cases expected at least 4 will not be damaged? 

ADDITIONAL TOPICS IN PROBABILITY 

11. The Poisson Law. In the problem of repeated trials it may happen 
that p is too small to permit the use of the normal approximation even 
though n is large. A different approximation, which is called the Poisson 
law or the law of small numbers , is now to be obtained for this case. 




ADDITIONAL TOPICS IN PKOBABILITY 


655 


SBC* 11] 

Starting with the formula for the probability of r successes in n trials, 

B(r) p r (l — p) w ~ r , 

r\(n ~~ r): 

we replace n! and (n — r)! by their Stirling approximations to obtain 


B(r) 


n v e 


-"V2™ 


r1(n 


r ) n ~ T e-(n~ T '>y/2r{n - r) 


V T { 1 “ P)"‘ 


n e 


r![l - (r/n)]"- r+> * 


- *)*-'• 


(11-1) 


Since the expected value of r is np, we can assume that r is small compared 
with n. In this case 1 


(-0 


Similarly, since p is small, 

(1 - p)"“ r ^ (1 - p) n - [(1 - ^ e“ np . 

Substituting these two expressions into (11-1) yields the desired law of 
small numbers: 

(np) r __ 

£(r) ~ c np , n large, np moderate. (11-2) 

r! 

The result may be written 

B(r) s ~ e"*, (11-3) 

r! 


where m = np is the expected number of successes. 

An application of this law to some specific cases may prove interesting. 
Suppose it is known that, on the average, in a large city 2 persons die 
daily of tuberculosis. What is the probability that x persons will die on 
any day? In this case the expected number of deaths is /i = 2, so that 

2 * 

B{x) =-e~ 2 . 
x\ 

1 The reader is reminded that lim (1 -f h) lfh ■» e as h approaches zero through posi- 

tive or negative values. See I. S. Sokolnikoff, “Advanced Calculus,” pp. 28-31, McGraw- 
Hill Book Company, Inc., New York, 1939. 



656 PROBABILITY 

Therefore we have the following table: 


[crap. 8 


— 

X 

B{x) 

X 

B(x) 

X 

B{x) 

0 

0.135 

2 

0.271 

4 

0.090 

1 

0.271 

3 

0.180 

5 

0.036 


The Poisson law has a significance far beyond its connection with the 
binomial distribution, as will now be shown. Suppose points x t are dis- 
tributed at random on the x axis in such a fashion that the following 
assumptions are valid : 

1. The probability that a given number of points is in a given interval 
depends only on the length of that interval (and not on any information 
we may have about the points in adjacent intervals). 

2. If P( Ax) is the probability of 2 or more points in an interval of 
length Ax, then P( Ax)/ Ax —» 0 as Ax 0. 

3. If Pi(Ax) is the probability of 1 point in an interval of length Ax, 
then P\{ Ax)/ Ax — ► k, a constant, as Ax 0. 

In tins case the probability P n (x) of w points in an interval of length x 
satisfies the Poisson law 

, (^) n 

P n (x) = (11-4) 

n! 


To prove this result, consider an interval (0, x 4- Ax) of length x -f Ax. We can 
have n points in this interval in three mutually exclusive ways. Either there are n 
points in x and none in Ax, or there are n — I in x and 1 in Ax, or there are fewer than 
n — 1 in z and at least 2 in Ax. The probability of this last alternative may be written 
c Ax, where « — ► 0 with Ax, in view of assumption 2. 

Thus, by total and compound probability, 

P*(x -f Ax) « P»(x)P 0 (Ax) 4 - P n „ 1 (x)Pi(Ax) 4- « Ax. 


Subtracting P n (x) from both sides and dividing by Ax give 


Pn(x 4- Ax) - F n (x) 
Ax 


Pn(x) 


Po(Ax) — 1 
Ax 


4“ P n~l(x) 


P i(Ax) 
Ax 


+ «. 


(11-5) 


Since there must be no point, 1 point, or more than 1 point in an interval of length 
Ax, We have 

Po(Ax) 4- Pi(Ax) 4- P( Ax) - 1 

which gives 

Po(Ax) - 1 Pi(Ax) P(Ax) 


Ax 


Ax 


Ax 


( 11 - 6 ) 



SEC* 11} ADDITIONAL TOPICS IN PROBABILITY 657 

Taking the limit 0, we obtain —k in (11-6), and hence taking the limit in (11-6) 
gives 

£ PJx) - -kP n (x) + fcP„-i(x), n > 1. (11-7) 

For n « 0 the term P n ~i(x) is to be replaced by zero, so that 

~ Po(x) - -kPo(x). 
ax 

This separable differential equation yields 

P 0 (x) « ce~ kx m e ~ k * 

where the constant c « 1 since Po(0) ** 1 ; that is, an interval of zero length is sure to 
contain no points. (This follows from assumption 2.) 

Substituting Po(x) in the relation (1 1-7) for n **■ 1 we get 

£ Pi{x) - -fcP,(x) + ke~ k * 

which yields P\(x) * e~**(fcr). Proceeding step by step or using mathematical induc- 
tion, wc obtain (11-4). 

The following are some of the phenomena which satisfy the assumptions 
1 to 3 quite accurately and which, accordingly, obey a Poisson law: the 
distribution of automobiles on a highway, the distribution of starting 
times for telephone calls, the clicks of a Geiger counter, the arrival times 
for customers at a theater ticket office. The first example is a spatial 
distribution, while the last three refer to distributions in time. 


Example 1. What is the probability that the ace of spades will be drawn from a deck 
of cards at least once in 104 consecutive trials? 

This problem can be solved with the aid of the exact law (7-4) as follows: The proba- 
bility that the ace will not be drawn in the 104 trials is 

B( 0) « mCo(H2)\ 5 K2) m - 0.133 

and the probability that the ace will be drawn at least once is 1 — 0.133 «* 0.867. On the 
other hand, Poisson’s law (11-2) gives for the probability of failure to draw the ace 

B(0) - „ e -s, 

0! 


Hence, the probability of drawing at least one ace of spades is 1 — e~ 2 ** 0.865. 

Example 2. Show that the constant k in the Poisson law (11-4) represents the ex- 
pected number of points in a unit interval. 

Since the probability of n points in a unit interval is 


Pn(l) 



OO 


E(n) - £ e~ k 

n— 1 


— 

n! 


n 


- e~ k k £ 


k n-i 

(n - 1)1 


- - k. 


the expected number is 



058 


PROBABILITY 


[CHAP. 8 


PROBLEMS 

1. By use of the Poisson law compute the probability of (a) just one ace in 6 tosses of 
a die, (b) just one double ace in 36 tosses of a pair of dice. Compare the binomial law 
for cases (a) and (5). Which of the two cases satisfies the assumptions of the text more 
exactly? 

2. The probability is 0.0025 that a nail chosen at random from the output of a cer- 
tain machine will be defective. What is the probability that a keg of 1,000 nails made 
by the machine will have at most 3 defective nails? Hint: The keg has "at most 3” if 
it has 0, 1, 2, or 3 exactly. Use the Poisson approximation. 

3. In Prob. 2 it is desired to have a probability of at least 0.95 that the keg has at 
least 1,000 good nails. How many nails should the manufacturer put into the keg? 
Hint: If he puts in n « 1,000 4- m nails, he wants a probability 0 95 that the number 
of defective nails will be at most m. Use the Poisson law, taking np ™ l,000p ** 2.5. 

4. On a certain one-way highway it is proposed to install a traffic signal which has a 
60-sec red interval but a long green interval. The speed of the cars may be taken as 
30 mph, and the expected number is 10 cars per mile of highway. Neglecting any 
effects of slowing down, find the probability that just n cars will be obliged to stop 
when the light is red. What is the probability that at most 5 cars must stop? What 
is the expected number that must stop? Hint Assume that the cars are distributed 
according to the law (11-4), and see Example 2. 

6. A certain circuit can transmit 3 telephone calls simultaneously The expected 
number of incoming calls is 1 per minute, and each call lasts 3 min. What is the proba- 
bility of getting a busy signal? Hint: You will find the line busy if 3 calls or more have 
come in during the preceding 3-min interval. Use (11-4). 

12. The Theory of Errors. In this section the methods of probability 
are used to analyze the effect of experimental errors in measurement. If 
n independent measurements give the values m u m 2> . . ., m M , we consider 
questions such as the following: What is the best estimate for the quantity 
being measured as determined by these measurements? What is the 
probability that this best estimate is within 1 per cent, say, of the true 
value? How much added precision is gained by increasing the number 
of measurements? 

Proceeding to the first question, let mi and m 2 be two independent 
measurements of an unknown quantity m (such as the mass of an electron, 
for instance). It is desired to find a best estimate for m based on the 
measurements mi and m 2 . To this end we denote the best estimate by 
0(mi,m 2 ) and seek to determine the function 0. Now, if both measure- 
ments are increased by a given amount a, it seems reasonable to assume 
that the estimate also increases by the amount a. In symbols, 

0(mi + a y m 2 + a) * 0(mi,m 2 ) + a. (12-1) 

This relation is now postulated. 

Similarly, if mi and m 2 are multiplied by a fixed quantity it is reason- 
able to suppose that the best estimate is likewise multiplied by j8. This 
requirement leads to 



BEC. 12] 


ADDITIONAL TOPICS IN PROBABILITY 


659 

B(Pm h Pm 2 ) « pe(m h m 2 ) ( 12 - 2 ) 

which is also postulated. [Equation (12-2) is quite obvious when we con- 
sider the effect of a change of units. For instance if grams are used in- 
stead of kilograms as the unit of mass, we expect the estimate in grams to 
be 1,000 times as great as the estimate in kilograms,] 

Finally, since the two experiments are carried out under substantially 
identical conditions, it does not matter which experimental result is m x 
and which is m 2 . We are thus led to postulate that 0 is symmetric: 


0(rai,m 2 ) = 0(m 2} mi). 


(12-3) 


It is a remarkable fact that the best estimate is wholly determined by 
these requirements; if 6 satisfies (12-1) to (12-3), then 0 must be the arithmetic 
mean , 


0(mi,m 2 ) 


mi + m 2 
* 2 


( 12 - 4 ) 


To establish (12-4), regard mi and m 2 as fixed and choose a — —m 2 in (12-1). There 
results 

0(mi,m 2 ) =* m 2 + 0(mi — m 2 , 0). (12-5) 

If this expression for 0(mi,m 2 ) is used in the left-hand member of (12-2), one obtains 

/5m 2 + 0(/3mi - 0 m 2 , 0) = 00(mi,m 2 ). (12-6) 

Whenever mi 5 * m 2 , the choice /3 ~ l/(mi — m 2 ) in (12-6) gives 

m 2 H- 0(l,O)(mi — m 2 ) ~ 0(mi,tn%) (12-7) 

if we multiply through by mi — m 2 And now (12-3) leads to 

m 2 -f 0(l,O)(mi — m 2 ) — mi + 0(l,O)(m* — mi) 

which implies 0(1,0) « Y 2 , Hence (12-7) yields (12-4). The case mi m 2 is even sim- 
pler; specifically, Eq. (12-5) gives 

d(m h mi) ** mi + 0(0,0) (12-8) 

and the choice /9 « 0 in (12-2) shows that 0(0,0) =» 0. 

By analogy with (12-4), one generally assumes that the best value for 
three or more measurements is also the arithmetic mean. Thus, 


mi + m 2 + m 3 
0(m 1 ,m 2 ,m 8 ) * ~ 


(12-9) 


We shall now use this assumption to determine the underlying probability 
distribution for the errors of measurement. 

Let the true value of the quantity being measured be denoted by t;. 
The errors, then, are 

x % ~ mi — v . ( 12 - 10 ) 



660 


PROBABILITY 


[chap. 8 

Since the experimental determinations are made under substantially 
identical conditions, these random variables are all assumed to have the 
same probability density f(x). And since the experiments are supposed 
to be independent, the joint density for two or three variables is given 
by the product: 1 

f(x u x 2 ) «/Cri)/(* 2 ) (12-11) 

f(x i,x 2 ,x 3 ) « f(xi)f(x 2 )f(x s ). (12-1 la) 

Our task is to determine the function f{x). 

Now, v is the true value of the quantity being measured. It is not a 
random variable, and it is not at the disposal of the experimenter. Never- 
theless, one can contemplate the effect of a change in v, and in particular, 
one can consider that value of v which would maximize the probability of 
the observed result. We now postulate that the value of v which maximizes 
this probability is the arithmetic mean of the measurements. In other 
words, the best estimate , (12-4) and (12-9), is assumed to be also a maximum- 
likelihood estimate. It will be found that this assumption 2 enables us to 
determine the form of the function / without any knowledge of the experi- 
mental process. 

If the probability (12-11) is maximum when 

v - -i-— A (12-12) 


then the logarithm of the probability is also maximum. Thus 

log /(mi -v) + log/(m 2 - v) + log/(m 3 - v) (12-13) 

is maximum, as a function of v, when (12-12) holds. Setting the derivative 
with respect to v equal to zero in (12-13), we obtain 

/'(?% - v) f'(m 2 - v) /'(m 3 - v) = q 

f(mi - v) f(m 2 - v ) f(m a - v) 


If F is defined by 



1 If we think of the errors as being discrete variables with / the frequency function, 
(12-11) is simply the law of compound probability for independent events. That is, 
the probability of making an error xi in the first experiment and X 2 in the second is the 
product of the individual probabilities. The corresponding result for continuous varia- 
bles and densities (stated in Sec. 6) is also a consequence of the theorem of compound 
probability. The notion of independence is discussed further in Sec. 13. 

* We shall suppose also that / is positive and twice differentiable, though these require- 
ments could be somewhat relaxed. 



SlffiC. 12] ADDITIONAL TOPICS IN PROBABILITY 661 

the foregoing result, in the notation (12-10), is 

F{x i) + F(x 2) + F(x 3) *» 0. (12-14) 

Equation (12-10) shows that (12-12) is equivalent to 

xi + x 2 + ~ 0. (12-15) 

Thus, (12-14) bolds whenever (12-15) holds. The corresponding statement 
for two variables, obtained from (12-11), is that 

F(x 1) + F(x 2 ) - 0 (12-16) 

whenever X\ + x 2 = 0, and for one variable, we have 

F(xi) = 0 when x x « 0. (,12-17) 

From (12-16) we get —F(x 3 ) ~ F(-~x 3 ) by choosing = X3, x 2 - — x 3 , 


and hence (12-14) gives 

F(x l ) 4- F(x 2 ) = ~F(x 3 ) » F(-x 8 ). 

Since —x 3 = X\ + x 2 by (12-15), the function F satisfies 
F(x x) + F(x 2 ) = F(xx + :r 2 ). 

Differentiating partially with respect to Xi and x 2 leads to 

F'(x 1) » F f (x 1 + x 2 ) and F'(x 2 ) = F'fci + x 2 ). 

Hence F'(x 1) = F'(x 2 ). Holding x 2 constant, we see that F'(x 1) is constant: 

F'(s,) « c 


and hence F(xi) 


cx i, since (12-17) gives F(0) — 0. 
/'(*) 


fix) 


F(x) « cx 


The relation 


yields /(x) = Ke Hcx * 

where the constant X may be found from 

l ~ f f(x) dx — K f dx. 

J — co • —00 


Since the integral diverges if c > 0, we set c =■ —2ft 2 to obtain 


1 

K 




by (10-1). Hence K 


h/y/r, and 


/(*) 


Vi 




-\/2jr 

V2ft 


(12-18) 



662 


PROBABILITY 


(CHAP. 8 


This result, known as the Gaussian law of error , states that the variable 
y/2 hx is normally distributed. Specifically, the probability of 

h < \/2 hx < t 2 (12-19) 

h ) h 

/ ^ 

Jiu 


IS 


~h*x* 


dx 


( 12 - 20 ) 


h /(V2M y/ir 

by (12-18), and the change of variable t = y/2hx shows that (12-20) is 


-4= /'• = s>(; 2 ) - «>«!). (12-21) 

v 2 tt 

The most important consideration justifying the use of this analysis in 
practice is that systematic errors must be eliminated. 

The constant h measures the accuracy of the observer and is known as 
the precision constant . That particular error which has probability Y to 
be exceeded in magnitude is called the probable error; it is found to be 
0.4769//t by use of (12-19), (12-21), and Appendix D. Another interpreta- 
tion of the constant h is afforded by considering the mean-absolute error 


r* 2h r<*> , , 1 0.5642 

25(1*1) = / |*l/(*) dx « ~~ I xe- h * xl dx = — — - = — — (12-22) 

■'—00 \/ 7T fl\/' w h 

and still a third interpretation is given by the mean-square error 

/ <*> 2h /■<*> „ , 1 

x 2 f(x) dx = — / arV* v dx - — -• (12-23) 

-00 V 7T •'O 2 h 4 

The final question mentioned at the beginning of this section concerns 
the effect of increasing the number of measurements n. Since x % = m x — v, 
we have 

x ~ ffl — v 


where the bar denotes the arithmetic mean: 


x = - 2x„ rn = - 2m,*. 
n n 

Thus, the error in the mean is the mean of the errors. It is likely to be 
smaller than the error in a single measurement because positive and nega- 
tive errors tend to cancel when we form 2x,. For the Gaussian distribution 
(12-18) the situation is especially simple; namely, x has a Gaussian distribu- 
tion with precision constant hy/n, whenever the independent measurements 
Xi have Gaussian distributions with precision constant h. Thus, if the in- 
equality | x | < a has probability p, then the inequality \$ | < a/y/n has 
the same probability p, This result shows how much more precision is 
attained by increasing the number of measurements. 



ADDITIONAL TOPICS IN PROBABILITY 


SBC. 13] 

The proof is omitted because it involves a tedious evaluation of multiple integrals.' 
However, the essential meaning of the result is that the “scatter” or “spread” for £ is 
1 /\/n times as great as the corresponding spread for x. When interpreted in this fashion 
the property follows from the results established in Sec. 14. 


PROBLEMS 

1. (a) Show that the sum of the squares of the errors 2(m t — v) 7 is least if the true 
value v happens to be the arithmetic mean of the measurements m,. ( b ) Deduce that 
the arithmetic mean m is a maximum-likelihood estimate for v when there are n inde- 
pendent measurements each satisfying (12-18). Hint: It is required to choose v so that 

/(*!,** X B ) - /(xO/fe) .../(X») - 

is maximum. Use the result (a). 

2. In a certain experiment which satisfies the conditions of the text, the probable 
error is 0.01. A measurement mi is about to be made. What is the probability that 
the interval (mi — 0.02, mi -f 0.02) will contain the true value v ? Hint: First find A, 
then note that the stated result happens if, and only if, \x\ \ < 0.02. 

13. Variance, Covariance, and Correlation. Two random variables 
x and y are said to be independent if the event x — x t and the event y = y 3 
are independent events for each choice of r t in the range of x and each y, 
in the range of y. In other words, knowledge that y has a particular value 
must not influence the probabilities associated with x. The numbers 
shown on two successive tosses of a die are independent in this sense 
(and so were the measurements m t considered in the last section). On 
the other hand, the number of heads in the first three tosses and in the 
first four tosses of a coin are dependent variables. 

The product xy of two random variables is a random variable which 
equals x l y } when x — x t and y = y r Although it is not usually true that 
the expectation of a product is the product of the expectations, this is the 
case when the variables are independent. In symbols, 

E(xy) « E(x)E(y), x , y independent. (13-1) 

The proof is simple. If pi is the probability that x = x t1 and if q is the 
probability that y = y J} then the assumed independence gives p^j for the 
probability that simultaneously x = x, and y = y v Hence 

E(xy) = ZZpiQjXiyj = (Sp,x,)(2^) = E(x)E(y). 

* See J. V. Uspensky, “Introduction to Mathematical Probability,” chap. 13, McGraw- 
Hill Book Company, Inc., 1937, for a direct verification. An indirect method based on 
the theory of moments is given in P. G. Hoel, “Introduction to Mathematical Statistics,” 
sec. 6.4, John Wiley <fc Sons, Inc., New York, 1954. See also M. E. Munroe, “The 
Theory of Probability,” pp. 91-96, McGraw-Hill Book Company, Inc., New York, 1951, 



664 PROBABILITY [CHAP. 8 

When a discussion involves several variables x, y, . . . , it is convenient 
to denote expectations by the letter m, with a subscript to indicate the 
variable. Thus, we write 

E(x) « Mx, E(y) = fiy 

and so on. For example, (13-1) in this notation takes the form 

Mxy ® MxMy, x, y independent. (13-2) 

To measure the deviation of a variable from its expected value y, one 
introduces a quantity a defined by 1 

a = Ve(x - m) 2 or a 2 = E(x - ju) 2 . (13-3) 

The expression a is called the standard deviation , and its square a 2 is called 
the variance . As for here, too, it is customary to use a subscript when 
several variables have to be distinguished. For example, 

al = E (: r - Mi) 2 ) 4 = FAy - m„) 2 - 

To illustrato the calculation of a variance by means of the definition, let x denote the 
number of heads obtained when 3 coins are tossed. Since n - E{x) — ^ we have the 
following table: 


x 

0 

1 

i 

2 i 

3 

X - fl = 

-H 

-H 

H 

H 

(x - - 

% 

H 

H 

% 

Probability « 

Ys 

Ys 

H 

K 


The definition of expectation now gives 

* 2 * B(X - M) 2 - Vs'H 4- H-H + H-H + H H » 

If £(x) = nx and E(y) ~ My, the quantity 

^xy “ E(x Mx)(2/ My) (13-4) 

is called the covariance of # and y. The covariance is a generalization of 
the variance, in that the special case y ~ x gives 

<4 - «(* - Mx)(* - Mx) - E(x - Mx) 2 - crj. 

As an illustration, let us compute when x is the number of heads obt&inod on the 
first 2 tosses and y the number obtained altogether in 3 tosses of an unbiased coin. 

1 The intent is l£f(® — m) 2 ], not [E{x — m)J*. 



SEC. 13] ADDITIONAL TOPICS IN PROBABILITY 665 

Here /»« >« 1, % % so that we have the following table: 


Event 

HHH 

HHT 

HTH 

HTT 

THH 

THT 

TTH 

TTT 

X ~~ Hz 

1 

1 

0 

0 

0 

0 

-1 

-1 

V - Mv 

X 

X 

X 

-A 

X 

-X 

-X 

-X 

Product 

H 

V 

0 

0 

0 

0 

V 

X 


Since the associated probabilities are }/%, we take times the sum of the entries in the 
last row to get 

<4 - H. (13-5) 

We shall now obtain an expression for a xy which is often more useful 
than (13-4). Expanding the product in (13-4) gives 

aly = E(xy — yn x — xy y + n x n y ) 

= E(xy) — E(y)n x - E(x)y v + 

Upon recalling that E(x) = y x and E(y) « y. y we get 

°iv = E(xy) - E(x)E(y) * M;ey - Ma;My , (13-6) 

which is the required formula. 

To apply this formula to the preceding example, we construct the following table: 


Event . 

HHH 

HHT 

HTH 

HTT 

THH 

THT 

TTH 

TTT 

x 

2 

2 

1 

1 

1 i 

1 



y 

3 

2 

2 

1 

2 

1 

1 

0 

xy 

6 

4 

2 

1 

2 

1 

0 



Taking % times the sum of the last entries gives E(xy) «* 2, and hence by (13-6) 

<4 - 2 - ( 1 )(«) * Yi. 

The special case x = y in (13-6) gives an alternative form 1 of (13-3), 
namely, 

cr 2 - E(x 2 ) - m 2 = «(**) - lE(x)] 2 . (13-7) 

As an illustration the reader may apply this formula to the preceding 

example to obtain 

4 - X - O) 2 - X, 4 “ 3 - (%) 2 - (13-8) 

1 Note that <r 2 gives the moment of inertia of the area under the distribution curve 

y =* f{x) about the line x ** y which passes through the center of mass. From this 
viewpoint (13-7) is the familiar formula for moment of inertia after a change of rotational 
axes. 














f 


666 


PROBABILITY 


{chap, 8 


If the variables x and y are independent, (13-2) and (13-6) give a xv « 0. 
Hence when <r xy & 0, the variables must be related. A quantitative 
measure of the strength of the relationship is given by the correlation 
coefficient p: 

, - A. (1W» 

VxOy 

For example, in the foregoing illustration (13-5) and (13-8) yield 


M _ 
VK vh 


ivt 


0.816. 


(13-10) 


Thus, if two variables x and y have a correlation coefficient p = 0.8, then 
they are about as strongly related as are the numbers of heads on the first 
two tosses and on the first three tosses of an unbiased coin. 


The correlation coefficient has the value 1 if y « x, and, as we have already observed, 
p ** 0 when x and y are unrelated. Moreover, p does not change if x and y are each 
multiplied by a constant factor. Thus, if the correlation coefficient indicates a certain 
strength of relationship for x and y, it will give the same strength of relationship for 2x 
and 3 y. Similarly, p is unaffected by addition of a constant; for instance, x — 2 and 
y — 3 have the same p as x and y. 

In spite of having these desirable properties, p is not always a reliable measure of 
dependence, and many statistical studies have led to erroneous conclusions through an 
incorrect interpretation of correlation. It is quite possible to have the variables so 
strongly related that y is a function of x and yet p * 0. Before a correlation coefficient 
can be used with confidence, one must know something about the underlying probability 
distribution. 

The variables x and y are said to have a bivariate normal distribution when 
f(x,y) m eW+itev+cit+dx+cv+f), conBt 


In this important case the theory of correlation has been fully developed, and it is 
found 1 that p actually does measure the strength of the relationship between x and y. 

Example: A variable x is said to be “normally distributed with mean a and variance 
tr 2 " when its density function is 

i _ y 

/(x) ** —y=— t 2 ^ 0 ' , p, <r const. 

V 2ir a 


Show that the mean is indeed y and the variance <r 2 . 
By the definition of expectation, 




*<*dt 


when we set t — (x — y)/e. Hence E{x - y) « 0, which gives E(x) - p. 
change of variable leads to 


E(x 


- „) J - -^L j°° Pe-H* dt - 


-0 

The same 


1 See Hoel, op, cU. t chap. 8. 



ADDITIONAL TOPICS IN PROBABILITY 


SNC. 14] 


667 


aa we see upon integrating by parte and using (10-1). Since j u «■ #(x), the latter result 
E(x ~ p) 2 is the variance by definition. 

Choosing ^ 0, <r « l/(A\/2), we obtain (12-18), and hence the precision constant h 

is given by 

*-vb as -“> 


[cf. (12-23)]. This fact gives a method for estimating h from the data, as we shall see 
in the following sections. 


PROBLEMS 

1. Compute <r 2 if x is uniformly distributed on the interval 0 < x < 1. 

2. Let x be the number on top and y the number on the bottom in a toss of a true 
die. Compute E(x), E(y), E(xy) f and the covariance. Does your work indicate that 
the variables are dependent? Find the correlation coefficient. 

8. Three coins are tossed. Let x be the number of heads shown by the first coin, 
whereas y is the number of heads shown by all the coins. Compute the correlation 
coefficient. Your result should be smaller than the value (13-10) Why? 

14. Arithmetic Means. In many applications one does not consider a 
single variable, but rather one obtains the mean of a large number of 
variables. For instance, if x is a measure of the length of a rod, one would 
make several measurements x { , r 2 , . . ., x n and use the arithmetic mean, 

_ *1 + * r 2 H h * r n , , v 

(14* i) 

n 

in accordance with the procedure of Sec 12. Here the .r,s are not the 
different values of a single variable but are n random variables describing 
the result of n independent measurements. 

Just as one uses <t x to indicate the standard deviation of the variable z, 
it is customary to let cr f denote the standard deviation of x. The following 
theorem enables us to compute <r f from a r x in many cases: 

Theorem. If the variables x t are independent, if they have the same ex- 
pectation E(x t ) ~ p and the same variance a 2 , then 

- - ik (14 - 2) 

For proof, observe that 

E(x x 4 — * + x n ) = E(x i) 4 — *4- E(x n ) = np. 

The variance of x x 4 b x n is therefore 

E(x i 4 b x n - np) 2 > 

which may be written 

E[(Xi — p) 4~ tea ~ p) 4 1- ten — m)] 2 . 



PKOBABILITT 


[chap. 8 


Expanding the bracket we obtain 

E [22 (*> ~ m) 2 + ]£ (*< - »){xj - p) J • (14-3) 

Since the variables are independent, the covariance of x, and Xj is zero 
for * j; that is, 

B(xi - n)(xj - n) » 0 . 

Also the definition of <r* gives 

a\ = E(Xi — m ) 2 - 
Hence, taking the expectation in (14-3) yields 

E(x j 4 f x„ — nil) 2 = n<r 2 . 

Dividing by n 2 we have 

r^ L± _ ±i _ 

L n J n 

which gives (14-2) upon taking the square root. 

The intuitive meaning of this result is approximately as follows: Suppose 
a single measurement varies over an interval of length / about the true 
value, so that l measures the scatter or spread. Then the mean of n in- 
dependent measurements will have a spread of the order of 1/ y/ n about 
the true value. 

To illustrate the use of (14-2) let x x = 1 if there is success at the zth 
trial in a set of independent trials with probability p> and let x x = 0 other- 
wise. For each variable x x we have x? — x x and hence 

E(xf) « E(x l ) ~pl + q<0~p. 

By (13-7) the corresponding variance is 

<rl = p - p 2 = p(l - p) = pq 

and (14-2) now gives 



For the variables x x considered in the foregoing paragraph the mean $ is simply the 
relative frequency tn/n, where m is the number of successes. We have, then, 

which shows again that the relative frequency m/n is likely to be close to p when n is 
large. The corresponding result for a general variable * is based on ( 14 - 2 ); it leads to 



ADDITIONAL TOPICS IN PROBABILITY 


SRC. IS] 

assertions concerning |l£(ap) — f | which are similar to the theorem established hi Sec. 
10 but of greater scope. 

Multiplying (14-4) through by « we get 

[E(m — np) 2 ]** • y/rvpq. 

This gives an interpretation for the quantity ■%/ npq that arose in connection with the 
normal law (Sec. 10) ; namely, \/npq is the standard deviation of the number of successes m. 

15. Estimation of the Variance. If x u x 2 , . . x n are n independent 
observations of a variable x, the sample variance is defined by 

s 2 - ~ - *) 2 - (Xi - £) 2 . (15-1) 

n 

Unlike the theoretical variance cr 2 , the sample variance is computed from 
the observations, hence is actually available. It will be seen, now, that 
s 2 can be used to estimate cr 2 . 

W.e have 

E(ns 2 ) - 2E(xi - x) 2 

- m(xi - M> - - m )] 2 

= 2[E(xi - ti) 2 - 2 E(xi - h)(x-m) + E(£ - M ) 2 ]. (15-2) 

Now, E(xi — M ) 2 — <r 2 by definition, and E(x — /i) 2 = <r 2 /n by (14-2). 
For the middle term in (15-2) we get 

1 

E(xi - p)(£ ~ m) = ~ E(x t - m) ( a:i H b x n - pn) 

n 

= - E(x< - „)(*< - M +•••)= - 2?(x, - M ) 2 = - tr 2 
n n n 

when we note that the terms not written explicitly are independent of x». 
That is, for i ^ j , Eq. (13-1) gives 

E[(x* — p)xj] - E{xi — v)E(xj) - 0-m » 0. 

Substituting into (15-2) yields the important formula 

Bins 2 ) = (n - l)<r. (15-3) 

If (15-3) is divided by n, we get 

£[(^1^! - " tr 2 (15-4) 

71 

upon recalling (15-1). On the other hand the definition of <r s gives 


£[(*. -m) ! ] - a 2 . 


(15-5) 



870 


PROBABILITY 


[CHAP, 8 

It k not surprising that (15-4) gives a smaller value than (15*5), inasmuch as the choice 
ft -* £ is the value of a* that minimizes (15-5) (cf. Prob. 1, Sec. 12), The fact that (15-4) 
should be smaller than (15-5) is especially clear when there is only one measurement, z\. 
In this case (15-4) gives zero because x\ ~ £. 

The foregoing remarks indicate that s 2 is not a suitable estimate of o 2 ; 
it has a tendency to be too small. But if we divide (15-3) by n — 1 for 
n > 2, we get 

which gives the following theorem : 

Theorem. Let Xi,X2, . . . , x n be n independent observations of a variable x , 
with n > 2. If s 2 is the sample variance , then the quantity 

6 2 » — s 2 (15-6) 

n ~ 1 

is an unbiased estimate of a 2 . That is, E(d 2 ) - a 2 . 

To illustrate the use of the theorem, let 

mi = 12, m 2 " 8, vi i =» 13 

be three measurements of an unknown quantity whose true value is v . The errors in 
the measurement are X{ » m % — v, but since 

Xi — i « m t — v — ffi -f v « mi — m (15-7) 

we can compute a 2 without knowing v. By (15-1) and (15-7), 

ns 2 = £(x, - x) 2 » S(m,' — m) a . 

In this example fh 11, so that 

ns 2 * (l) 2 + (~3) 2 + (2) 2 * 14. 

Hence an estimate for <r 2 is 



According to (13-11) the precision constant h is estimated as h gg l/(\/2 6) = l/\/l4 
m 0.27. In statistics it is shown how one can determine the reliability of an estimate 
such as this, though we do not pursue the subject here. 1 

PROBLEMS 

1, A certain experiment gave the measurements 

m, - 17, 21, 20, 18, 14. 

Obtain an unbiased estimate for the variance of a single measurement, and from this, 
estimate the precision constant. 

1 See Hoel, op. cil. t chap. 10. 



1 


SEC. 15] ADDITIONAL. TOPICS IN PROBABILITY 67 1 

2. If the precision constant in Prob. 1 can be assumed exactly equal to your estimate 
of it, (a) what is the probability that the next measurement will be within 0.5 of the 
true value? ( b ) How many measurements must you make if you want a probability 
0.95 that the mean of those measurements will be within 0.1 of the true value? Hint: 
Use the fact that the precision constant of the mean is h\/n if that of a single measure- 
ment is h. 

3. In a certain measuring routine the cost of equipment and materials is negligible 
but the time required is proportional to the number of measurements. Give a rational 
method of adjusting the salaries of two observers whose working speeds are #i and b% if 
the precision constants of their measurements are h\ and h*. Hint: Consider the number 
of measurements each must make to attain equal reliability in the respective arithmetic 
means. 

4. Discuss Prob. 3 if the cost of equipment is proportional to the length of time it is 
used and the cost of material is proportional to the number of measurements. 




CHAPTER 9 


NUMERICAL ANALYSIS 




Solution of Equations 


1. Graphical Methods 677 

2. Simple Iterative Methods 679 

3. Newton's Method 684 

4. Systems of Linear Equations. The Gauss Reduction 687 

5. An Iterative Method for Systems of Linear Equations 689 

Interpolation. Empirical Formulas. Least Squares 

6. Differences 691 

7. Polynomial Representation of Data 694 

8. Newton's Interpolation Formulas 696 

9. Lagrange's Interpolation Formula 699 

10. Empirical Formulas 701 

1 1 . The Method of Least Squares 702 

12. Harmonic Analysis 711 

Numerical Integration of Differential Equations 

13. Numerical Integration 715 

14. Euler's Polygonal Curves 721 

15. The Adams Method 723 

16. Equations of Higher Order. Systems of Equations 727 

17. Boundary-value Problems 730 

18. Characteristic-value Problems 731 

19. Method of Finite Differences 734 


675 




The principal concern of numerical analysis is with the construction of 
effective methods for the calculation of unknowns entering in the formula- 
tion of a given problem. Since every formulation of a practical problem 
involves assumptions and approximations, it is senseless to seek unknowns 
to a higher precision than is warranted by the initial data. A simple and 
perhaps crude technique giving the desired values within specified limits 
of tolerance is always to be preferred to an involved method capable of 
yielding an arbitrary degree of accuracy. 

In recent years the growth of numerical analysis was accelerated by the 
demands of science and technology for numerical solutions of many pressing 
problems. High-speed computing machines produced for coping with 
such problems are certain to open new vistas in science and leave a pro- 
found imprint in all fields of human activity. 

It is the object of this chapter to present the rudiments of numerical 
analysis essential to all concerned with the processing of numerical data. 
Inasmuch as the understanding of principles must precede the acquisition 
of computing skills, the emphasis in the following sections is placed on 
basic ideas and general methods rather than on special techniques useful 
in solving this or that problem. Among topics included here are the 
determination of real roots of algebraic and transcendental equations, 
the basic method for solving systems of linear equations, the elements of 
interpolation theory, and its bearing on curve fitting and numerical so- 
lution of differential equations. 

SOLUTION OP EQUATIONS 

1. Graphical Methods. Geometric considerations usually are a useful 
guide in the construction of analytic methods of solution of practical 
problems. This is particularly true in the problem of determination of 
numerical values of the roots of algebraic and transcendental equations . 1 

1 A polynomial equation x n + aix n ~~ l H f a* » 0 is called an algebraic equation. 

An equation F(x) *= 0 which is not reducible to an algebraic equation is called Iramcen* 
dental . Thus, tans — x * 0 is a transcendental equation, and so is e x -f 2 coax *■ 0. 

677 



078 


NUMERICAL ANALYSIS 


[chap. 9 


If Fix) is a real continuous function, the equation 

Fix) - 0 (1-1) 

may have real roots. The approximate values of such roots can be de- 
termined by graphing the function y = F(x) and reading from the graph 
the values of * for which y = 0. This familiar procedure for graphical 
determination of real roots can frequently be simplified by rewriting (1-1) 
in the form 

m - g(x). (1-2) 

The abscissas of points of intersection of the curves y = fix) and y = g(x) 
will obviously be the roots of (1-2). 

Thus, an approximate value of the real root of 

Fix) m x 3 - 146.25* - 682.5 * 0 

can be found by graphing the function 

y = * 3 - 146.25* - 682.5. 

It is simpler, however, to plot the cubic 

y ** * 8 

and the straight line (Fig. 1) 

y « 146.25* + 682.5 

and read off from the graph the abscissa of their point of intersection P 0 . 

An obvious disadvantage of graphical 
methods is that they require plotting curves 
on a large scale when a high degree of accu- 
racy is desired. To avoid this, one obtains 
more precise values by applying one of the 
several methods of successive approxima- 
tions discussed in Secs. 2 and 3. All these 
methods require that the desired root be 
first isolated. That is, they call for the 
determination of an interval which contains 
just the root in question and no others. 
If Fix) is a continuous function, and if for 
a certain pair of real values * «* *i, * « * 2 , 
* the signs of F{x{) and F(*a) are opposite, 
then it is obvious that Fix) * 0 has at least 
one real root in the interval (*i,* 2 ). If there 
are several roots in (*i,* 2 ), one usually nar- 
rows down this interval by a succession of judicious trials until an interval 
is obtained which contains just the desired root. For efficient applies- 




SEC, 2] SOLUTION OF EQUATIONS 679 

tion of the successive-approximations methods it is desirable that this 
interval be as small as possible. 

We note in passing that no general methods are available for the exact 
determination of the roots of transcendental equations. Also, there are 
no algebraic formulas for the solution of general algebraic equations of 
degree higher than 4. The so-called Cardan and Ferrari solutions of the 
cubic and quartic equations require the calculation of cube roots of quan- 
tities which themselves are square roots. Generally it is simpler to obtain 
the desired approximations by methods described in the following sections 
than to make use of Cardan's formulas. 1 

PROBLEMS 

1. Find graphically, correct to one decimal, the real roots of : 

(a) 2* - x 2 - 0; (b) x* - x - 1 - 0; (c) x 5 - x ~ 0.5 - 0; (d) e* + x - 0; 
(c) tan t—z=0, v < x < 3tt/2. 

Isolate the roots (that is, for each root find an interval which contains just that root 
and no others). 

2. A sphere 2 ft in diameter is made of wood whose specific gravity is %. Find to 
one-decimal accuracy the depth h to which the sphere sinks m water. Hint: The volume 
of a spherical segment is vh\r — h/Z). The volume of the submerged segment is equal 
to the volume of displaced water, which must weigh as much as the sphere. If water 
weighs 62 5 lb per ft 3 , 

62.5 -I*-.?- 62.5. 

and since r ■» 1, we have h 3 — 3 h 2 + % » 0. 

2. Simple Iterative Methods. When real roots of Eq. (1-1) have been 
isolated, there are many methods for computing them to any degree of 
accuracy. These all depend on the application of some iterative formula 
which furnishes values of the succeeding approximations from the preced- 
ing ones. The nature of restrictions imposed on the function F(x) in the 
equation 

F(x) - 0 (2-1) 

in the two basic iterative methods discussed here is obvious from the 
description of the methods. The simplest of these is the method of linear 
interpolation , also known as the method of false position. 

Let the root x 0 of (2-1) be isolated between x x and x 2 - Then, in the 

1 A numerical determination of the roots of algebraic equations is frequently accom- 
plished by some method of synthetic division (such as Horner's method) or by the root- 
squaring method (Graeffe’s method). These special methods are discussed in many 
books. See, for example, F. B. Hildebrand, “Introduction to Numerical Analysis,” 
McGraw-Hill Book Company, Inc., New York, 3956. The methods of Secs. 2 and 3 
of this chapter apply to all types of equations and are generally adequate for the deter- 
mination of real roots. 



680 


NUMERICAL ANALYSIS 


{chap. 9 


interval (x h x 2 ), the graph of y #* F(x) may have the appearance shown 
in Fig, 2. If the points P\ and P 2 in Fig. 2 are joined by a straight line, 

it will cut the x axis at some point 
x 3, which usually is closer to the 
root x 0 than either X\ or x 2 . But 
from similar triangles, 



~F(*i) 


3*2 - *3 


F(x 2 ) 

and on solving for we get 
XiF(x 2 ) - x 2 F(x l ) 

^*3 


(2-2) 


(2-3) 


F(X 2) - F(.n) 

To obtain a (‘loser approximation 
to ,t 0 , v\e can determine the x inter- 
cept of the straight line joining the 
point P$ in Fig 2 with the point P 2 
and thus obtain the next approximation ,r 4 . By repeating this process ve 
obtain a sequence of values 


*r3> » « * ? Xn } 

which generally converges to rr () . The process described here is precisely 
that used in interpolating tabulated values of logarithms and other func- 
tions. In effect, it replaces a small portion of the curve by a straight line*. 
Another useful iterative method is based on rewriting (2-1) in the form 


Now, if the real roots of 


/« - gb). 
fix) = c 


(2-4) 


can be determined for every real c, we can proceed as follows. Let x x be 
an approximate value of the root x (J of (2-1). This, of course, is also an 
approximate root of (2-4), since (2-1) and (2-4) are equivalent equations 
On setting x *= xi in the right-hand member of (2-4) we get the equation 

f(x) = g(: ri), (2-5) 

which by hypothesis w r e can solve. If the solution of (2-5) is x 2 , we obtain, 
on setting x ~ x 2 in the right-hand member of (2-4), 

fix) « g(x 2 ). (2-6) 

The solution £3 of (2-6) we call the third approximation, and in general, 
^ the nth approximation x n is determined by solving 

f(x) * g{x n „ x ). 


(2-7) 



«BC. 2] SOLUTION OF EQUATIONS 681 

From the geometric interpretations of this procedure, which we give 
next, it will be seen that the sequence x X) x 2y * . . , x n , ... converges to the 
root x 0 of (2-1) if, in the interval of length 2\xx — x 0 j centered at x 0 , we 
have 

!/'(*) I > W(*)\ (2-8) 

and the derivatives are bounded. 

Suppose, first, that the slopes of the curves 

y * /(«), y - g(x) (2-9) 

in the interval (x 0 ,xi) (Fig. 3) have the same sign and satisfy (2-8). When 
x = xi is taken as the first approximation to x 0 , Eq. (2-5) yields the second 



approximation x 2r which corresponds to the abscissa of the point of inter- 
section P 2 of the straight line y = g(x x ) with y — /(x). Equation (2-6) 
gives J3, which is the abscissa of the point of intersection P3 of the straight 
line y = fif(x 2 ) with y = /(x), and so on. The sequence x Xt x 2l ... 
obviously converges to xq. 

The situation when the slopes of the curves (2-9) satisfy (2-8) but are 
opposite in sign is illustrated in Fig. 4. The value x 2 determined by 
solving (2-5) is the abscissa of the point of intersection P 2 of y « f(x) 
with y — ^(xi). It lies on the opposite side of the root from Xj. The 
third approximation x% is the abscissa of the intersection of y * g(x 2 ) 
with y = /(x), and it lies on the same side as xi but nearer to Xq. In Fig. 3 
the approach to the intersection Pq is along a staircase path, while in 
Fig. 4 it is along a spiral. In either case, the rapidity of convergence 1 
depends on the nature of the functions /(x) and g{x). 

1 Some criteria for the speed of convergence are given in Hildebrand, op. cii. 



682 


NUMERICAL ANALYSIS 


[chap. 9 



Example 1. Determine the approximate values of the real roots of 

e* — 4x m 0. (2-10) 

The real roots of this equation are the abscissas of the points of intersection of the 

curves y «■ e* and y « 4x shown in Fig. 5. 
It appears that the smaller of the roots, xo 
lies in the vicinity of x «* 0.3. The larger 
root, £o, is close to x ■» 2 1. Since for x * x 0 
the slope of y *■ 4x is greater than that of 
y ** e*, we write (2-10) in the form 

x =» 14e*, 

so that in the notation of Eq. (2-4) 

f(x) « x and tf(x) « 14«** 

The sequence of approximations x n according 
to (2-7) is thus determined from 

Xn 4 i - »- 1,2,.... (2-11) 

If we take x\ ® 0.3, we get 1 from (2-11) 

x 2 - he 0 ’ 3 » 14(1*34986) « 0.3374 

a* - 14 ^* « 34(1.40130) - 0.3503 

X4 « J 4 e^ «= 14(1.41949) - 0.3549 

x 6 « « 14(1.42603) * 0.3565 

x 6 - 14«* 5 - 14(1.42832) - 0.3571 

X1 m 14 ^* - 14(1.42917) « 0.3573. 

1 In performing these calculations it is convenient to use tables such as ‘ 'Table* of 
Exponential Functions/' National Bureau of Standards, Washington, D.C., 1951. 




683 


BBC. 2) SOLUTION OF EQUATIONS 

If only three-decimal-place accuracy is required, the computations can be terminated at 
this stage. 

To obtain the second root we note that at x <* $o, the slope of y • 4x is less than that 
of y *> e* If we write (2-10) in the form 

e* — 4x 


or x « log 4x, 

so that/(x) « x and g(x) » log 4x, then the condition (2-8) is satisfied at x ■» Bo. 

The desired sequence (x n ) is now given by 

x n +i - log 4x n , n * 1, 2, . . 

and we can take xi *» 2.1. 

Using tables of natural logarithms 1 we find 

X2 « log 4xi ** log 8.4 *» 2.12823 
x 8 «■ log 4x2 *» log 8.5129 » 2.14158 
X4 *> log 4 x 8 ** log 8.5663 =* 2.14783 
Xff ■ log 4x4 ** log 8.5913 2.15075 

x e - log 4x 6 - log 8.6030 ~ 2.15211 
x 7 « log 4xe - log 8.6084 - 2.15273 
x 8 - log 4x 7 * log 8.6109 - 2.15303 
x 9 m log 4x 8 - log 8.6121 * 2.15316. 

The value of the root $o, correct to three decimals, is 2.153. We do not give a dis- 
cussion of the errors in the approximations obtained by such calculations because a 
rigorous analysis of errors in the iterative procedures is fairly involved. 1 

Example 2. Find an approximate value of the real root of 


near x * 3r/2. 

From the graphs of 


x — tan x » 0 


y — x and y «• tan x 


( 2 - 12 ) 


in Fig. 6, it appears that Eq. (2-12) has just one real root in each of the intervals 

(2n — l)w/2 < t < (2n *f l)r/2, where n - 0, ±1, =fc2, 

It is convenient to rewrite (2-12) in the form 

x ■» tan” 1 x, 

so that in the notation (2-4) /(x) «■» x and g{x) — tan*" 1 x. This choice assures that the 
condition (2-8) is satisfied at the root xo. 

The sequence of approximations this time is given by 

Xn+i » tan~ l x n , n m 1, 2, . . 

1 For example, “Tables of Natural Logarithms/’ National Bureau of Standards, 
Washington, D.C., 1941. 

* A brief discussion is contained in Hildebrand, op. dt. } chap. 10. 



684 NUMERICAL ANALYSIS [CHAP, fl 

On taking t% ■» &r/2 «* 4.7124 radians, we find 

x 2 ~ tan" 1 4.7124 » 4.5033 
x% •» tan" 1 4.6033 * 4.4938 
X\ - tan- 1 4.4938 « 4.4935, 

which suggest that the root xo, correct to three decimals, is 4.493. 



These examples indicate that if it is possible to write Eq. (2*1) in the form 

x « g(z) t 

and if | g'(x) | < M < 1 in the interval of length 2 1 z\ — xo | centered at xo, then the 
recursion formula giving the desired approximating sequence is 

x n +i - gr(xn), n * 1, 2, (2-13) 

PROBLEMS 

1. Use both methods of this section to obtain, correct to two decimals, the values of 
the real roots in Probs. 1 and 2 of Sec, 1. 

2. Find in the manner of the examples of this section the real roots of x B — z — 0.2 ® 0 
correct to three decimals. 

3. Newton’s Method* The successive terms in the approximating se- 
quence in the method of false position (see Fig. 2) are determined by the 
intersection of the secant line with the x axis. Newton proposed con- 
structing an approximating sequence determined by the intersection with 
the x axis of the tangent line to the curve y « F(x ). 

Thus let the root x = x 0 of 


F{x) - 0 


(3-1) 



SEC, 3} SOLUTION OF EQUATIONS 685 

lie in the vicinity of x ® x\ (see Fig, 7). The equation of the tangent 
line to y « F(x) at Pi (#1,2/1) is 


y - F(x 1) * F'(x 0(# ~ #1). (3-2) 


If the curve y = F(x) has the appearance shown in Fig. 7, the tangent 
line (3-2) cuts the a; axis at #2, which is a better approximation to the root 
than x\. To determine x 2 we set 


y =* 0 and find 


#2 = #1 ~ 


F(x 0 

F'(#i) 


if F'Ctj) 5^ 0. Having determined 
x 2 , we find in the same way that the 
tangent to y = F(x) at P2[#2,F(x2)] 
intersects the axis at 


#3 = #2 


F(r 2) 
>(#2)’ 


and in general, 


#«+i 



F(x n ) 

F'(#n) f 


1 , 2 , 


(3-3) 


The geometric considerations indicate that when y = F(x) is a mono- 
tone increasing or decreasing function in the interval (xi,.r 2 ) [so that 
F'U) does not change sign] and when there is no point of inflection in 
this intena) [so that F”(x) does not change sign], the sequence (3-3) 
converges to the root x 0 . 



The situations corresponding to the cases when there is a point of in- 
flection or a horizontal tangent to y = F(x) in the vicinity of the root are 
illustrated in Figs. 8 and 9. It is clear from these figures that in these 
cases the sequence (3-3) need not converge to xq. Thus, before applying 
Newton’s method one should examine the behavior of F'(x) and F"(x) 
in the vicinity of the root. 



686 NUMERICAL ANALYSIS [CHAP. 9 

Example; Find the angle subtended at the center of a circle by an arc whose length is 
double the length of its chord. 

Let the arc BCA (Fig. 10) be of length 2 BA. If the angle subtended by this arc at 
the center of the circle is 2x radians, then the arc BCA «® 2xr while BA m 2r sin x f 
r being the radius of the circle. 

Our problem requires that 

2 xr ** 4 r sin x, 

or x — 2 sin x » 0. (3-4) 

On graphing the functions y «■ x and y * 2 sin x (Fig. 11), we see that they intersect at 
x «* 0 and at x » 1.88 radians, approximately. We reject the trivial solution x « 0. 




Since y * x — 2 sin x is obviously monotone increasing and has no point of inflection 
near the root xo, we can apply formula (3-3) with n = 1.88. Wc find 

ii — 2 sin ti 

X 2 « x\ ~ 

1 -- 2 cos Xj 


The third approximation is 


1.88 - 2 sin 1.88 
1 - 2cos 1.88~ 


Xf 


X 2 — 2 sm X 2 
1—2 COS X2 


1.896. 


* 1.896 - 


1.896 —J2 sin 1 896 
1 - 2 cos L896~~ 


1.8955, 


which is nearly the same as xa. The angle subtended by the arc BCA , as given by this 
approximation, is 3.7910 radians. 


PROBLEMS 

1. Calculate by Newton's method the roots in Examples 1 and 2 of Sec. 2. 

2. Solve by Newton’s method Prob. 2, Sec. 1. 

$* Find to three decimals by Newton’s method the angle subtended at the center of 
a circle by a chord which cuts off a segment whose area is one-fourth that of the circle. 

4 . Find by Newton’s method to three decimal places the real roots of the following 
equations: (a) x - coex - 0; (6) x -f e* * 0; (c) x 4 ~ x - 1 « 0; (d) x 9 - 25 - 0; 
M as 1 — x - 0.2 «• 0. 



SOLUTION OF EQUATIONS 


687 


SEC, 4] 

4. Systems of Linear Equations. The Gauss Reduction. No doubt the 
reader is familiar with Cramer’s rule for solving systems of n linear equations 
in n unknowns by determinants. 1 Although Cramer’s rule is important 
in numerous theoretical considerations, it is of questionable practical value 
when the given system contains more than two unknowns. Usually it is 
easier to obtain solutions by some process of elimination of unknowns. 
The simplest practical method for solving systems of linear equations, 
based on the idea of elimination, is the Gauss reduction method. Its 
several variants form the basis for most techniques used in the solutions 
of large systems of equations. 2 

The idea of the method is simple. Let it be required to solve a system 
of n linear equations 

0ii*i + 0 1 2 * 2 H b a ln x n = Ci 

021*1 + «22*2 H h a 2 n*n = C 2 (4-1) 


0wl*l T* 0n2*2 “4* * * ' "4“ 0nn*n C n 


in n unknowns We divide the first equation in (4-1) by on, solve for 
a*i, and use the result to eliminate X\ in the other equations. The resulting 
system of n — 1 equations in rr 2 , . . x n is treated in the same way. 
That is, we divide the first of these equations by the coefficient of x 2 and 
use the result to eliminate x 2 from the remaining equations. After con- 
tinuing the process n times 3 we obtain an equivalent system 


*1 + a l2 x 2 + 013*3 + * • • + 0ln*n — 

*2 “4" 023*3 + * * * 4" 02n*n = C 2 


*n — 1 "4“ 0n- 


n*n — C n — i 
*n ~ 


(4-2) 


provided the given system has a unique solution. The substitution x n = c n 
in the preceding equation in the set (4-2) yields the value of x„_j, and 
by working backward we obtain in succession the values of x n _ 2 , x„_ 3 , . . . , 

*i. 

In practice the Gauss reduction can be performed in the manner indi- 
cated in the following example. 


1 A summary of the properties of determinants and Cramer's rule are given in Ap- 
pendix A. 

1 Among such variants are the Crout and the Gauss-Jordan reductions. These are 
described in Hildebrand, op. cit ., and in many other books. 

1 If the coefficient of x r in the rth equation vanishes, it is necessary to renumber the 
variables or equations. 



NUMERICAL ANALYSIS 


[CHAP. 9 


Example: Solve the system 

2.843xi - 1.326X5 4* 9.841a* - 5.643 

8.673x1 4- 1,295*2 - 3.215*8 - 3.124 (4-3) 

0.173*1 - 7.724*2 4- 2.832*3 * 1.694 


by the method of Gauss' reduction. 

On dividing each equation in (4-3) by the coefficients of %\ in that equation, we get 
*i - 0.46641*2 4- 3.4615*8 - 1.9849 

xi 4- 0.14931*2 - 0.37069*8 * 0.36020 (4-4) 

*i - 44.647*2 4- 16.370*8 - 9.7919. 

The subtraction of the second equation in (4-4) from the first and the third gives 
—0.61572*2 4- 3.8322*3 * 1.6247 
—44.796*2 4* 16.741*3 - 9.4317 
and, on dividing these by the coefficients of * 2 , we find 


*2 - 6.2239*3 = -2.6387 
*2 - 0.37372*8 * -0.21055. 


(4-5) 


Subtracting the second equation from the first in (4-5) yields 


-5.8502*3 - -2.4282, 
so that *s » 0.41506. 

The reduced system consists of the first equations in (4-4) and (4-5) and Eq. (4-6). 
*1 - 0.46641*2 4- 3.4615*3 - 1.9849 


(4-6) 
It is 


*2 - 6.2239* 8 * -2.6387 (4-7) 

*s « 0.41506. 

The substitution of the value of *3 from the last into the second equation of (4-7) gives 
*2 * -2.6387 4- 6.2239(0 41506) * -0.055408 
and the first reduced equation finally yields 

*1 * 1.9849 +0.46641 (-0.055408) - 3.4615(0.41506) - 0.52232. 

There are numerous modifications of the procedure just indicated, some of which are 
adapted for computations on desk calculators while others are more suitable for high- 
speed electronic computers. 


PROBLEM 


Use Cramer’s rule and also apply the Gauss reduction to solve the following systems; 

(a) 2* + y + 3« « 2, 

3* — 2 y — z » 1, 

* — y + 2 «* — 1 ; 

„ (&) 2*i + *2 + 3*g + *4 » —2, 

5*1 + 3*a — *8 — *4 1, 

*1 — 2*2 + 4*a + 3*4 ** 4, 

3*1 ^*2 +*j»2; 



SEC. 5] SOLUTION OF EQUATIONS ' 089 

(c) LS29»i + L415xi - 2.291X8 - 0.532, 

L395xi - 0.531x8 - 1.211, 
l.OOlx! + 2.093X8 - 0,556. 

5. An Iterative Method for Systems of Linear Equations* Except for 
the round-off errors the Gauss reduction method explained in the pre- 
ceding section is exact. When the determinant of the system (4-1) is 
different from zero, it yields the desired solution after a finite number of 
steps. However, successive steps leading to an equivalent triangular 
system (4-2) may prove laborious and ill-adapted to machine calculations. 
For this reason, a variety of iterative methods, which in theory require 
an infinite number of steps to obtain an exact solution, have been devised. 

One of these methods, due 1 to L. Seidel, is based on the use of the 
iterative formula (2-13). The convergence of any iterative method ob- 
viously depends on the character of the system under consideration. 

In many cases the system (4-1) can be rewritten so that in the zth 
equation the coefficient an of the unknown x» is numerically large compared 
with other coefficients. That is to say, the coefficients along the diagonal 
of the system (4-1) dominate the other coefficients. In this event by solving 
the fth equation for x» we can rewrite such a system (4-1) in the form 

1 

X\ = (Ci ~ aj2^2 — U 13 X 3 a Jn X n ), 

an 

1 

X 2 s= (c 2 ~ U2iXi ~ ^23^3 — • * * — a2 n X n )j (5-1) 

a22 


x n = (c n a*iXi a n2 x 2 * * * a n , n __ix n __j). 

a n n 

If we set Xi =* X 2 — • * • = x n ** 0 in the right-hand members of (5-1), we 
obtain 

X<‘> - * = 1, 2, (5-2) 

«u 

which is called the first approximation to the solution of (5-1). 

The substitution of this first approximation in the right-hand members 
of (5-1) yields the second approximation x (2 \ and so on. The cycle is 
then repeated with the expectation that the values x[ k) after the kth iteration 
are not substantially altered by further iterations. 2 

1 Generally called the Gauss-Seidel method. 

•There are several criteria for convergence of this process which generally are not 
easy to verify. It is known that when the coefficients in (4-1) are symmetric (so that 
<H } m aj t ) and the matrix (an) is positive definite, the Seidel process always converges. 
See Hildebrand, op. dt, , for a brief discussion of several criteria. 



NUMERICAL ANALYSIS 


690 


(chap. 9 


In practice the iteration process described above is usually modified by 
taking as the first approximation xi 1 * the value of xi obtained from the 
first equation in (5-1) by setting x 2 — x 3 * • * * » x n * 0. Using this 
value in the second equation in place of x x and setting x 3 ** x 4 =*•••» x n 
*# 0, one obtains the approximation x ( 2 l \ To obtain one inserts for 
x% and x$ the values x \ l) and xjj 1 * in the third equation and sets x 4 « xg — 
. . . == x n * 0. Finally, to get the value of x one uses previously found 
values x[ l \ . . ., x\~ x in the last equation of the system (5-1). This process 
is repeated to obtain approximations of higher orders. 

This particular choice of approximations usually improves the rapidity 
of convergence of the process. We illustrate it by an example. 

Example: The system (4-3) can be rewritten in the form 

8.673xi + 1.295x2 - 3.215x 3 - 3.124 

0.173xi - 7.724x2 4* 2.832x g - 1-694 (5-3) 

2.843xi - 1.326X2 4- 9.841x a - 5.643 

in which the diagonal coefficients dominate. 

We next write (5-3) in the form (5-1) and get 

*1 - (3.124 - 1.295^2 + 3.215*,) 

o.o73 

x, - - - 1 - (1.694 - 0.173x, - 2.832x,) (5-4) 

mmm i ♦ /Z4 

Xt - - (5.643 - 2.843X! 4- 1.326x2). 

9.841 

To obtain xi n we set x% ■» x 3 =* 0 in the first equation in (5-4) and find 

Xi 1 * - - 0.36020. 

8.673 


Inserting this value for xi and setting x 3 « 0 in the second equation in (5-4), we get 

xi l) - -0.21125. 

Finally, xj 1} *» 0.44089 is obtained by using the values xS 1} and in place of xi 
and X8 in the third of Eqs. (5-4). 

A repetition of the process yields second, third, and higher approximations. These 
are recorded in the table: 


k 

1 


3 

4 

5 

6 

7 

4*> 

0.36020 

0.56517 

0.51780 

0.52312 

0.52220 

0.52236 

0.52233 

*?> 

-0.21125 

-0.04523 

-0.05852 

-0.05501 

-0.05550 

-0.05543 

-0.05544 


0.44089 

0.40694 

0.41594 

0.41488 

0.41508 

0.41505 

0.41505 











SBC- 6 ] INTERPOLATION AND EMPIRICAL FORMULAS ' 691 

A comparison with the values found in Sec, 4 by the Gauss reduction method shows 
that in this problem six iterations were necessary to get four-decimal accuracy. 


INTERPOLATION. EMPIRICAL FORMULAS. LEAST SQUARES 

6. Differences. One of the problems connected with the analysis of 
experimental data concerns the representation of such data by analytic 
formulas. Thus, we may wish to represent, either exactly or approxi- 
mately, a set of observed values (x»,y t ) by some relationship of the form 
y *» f(x). In such analysis the concept of differences is important. 

We consider a set of pairs of values (; Xi t y % ), where i * 0, 1, . . n, which 
can be represented by points in the xy plane. The differences between 
successive pairs of ordinates yi+i and y> we call the first forward differences 
of the y& and we denote them by Ay*. Thus, 

= Ft+i - Vi, i - 0, 1, 2, , . ., n. (6-1) 

The second forward differences are defined by 

A 2 yi « kyi+i - A yi 

and, in general, the kth forward differences are 

A k y t = A h ~ l yi+i - A h ~ l yi. (6-2) 

These differences are usually represented in a tabular form: 


Table 1 





692 


NUMERICAL ANALYSIS 


[CHAP. 9 

in which the quantities in each column represent the differences between 
the quantities in the preceding column. These are usually placed midway 
between the quantities being subtracted, so that the forward differences 
with like subscripts lie along the diagonals indicated in the table by arrows. 

We note that if the rth differences A T y t are constant, then all differences 
of order higher than r are zero. 1 
Now, it follows from (6-1) and (6-2) that 

Vi m Vo + &Vo 

y* * y% + Ayi « (Vo + &Vo) 4- (A 2 y 0 + Ay 0 ) « yo + 2A y 0 + A 2 y 0 

V$ ** V% + &V 2 ~ (Vo 4* 2Ayo 4* & 2 yo) 4- (A 2 yi 4* AyO 

5=5 (yo 4“ 2Ayo 4~ A 2 yo) 4* (A 3 yo 4- A 2 yo 4- A 2 yo 4“ Ayo) 

J53 yo 4- 3 Ayo + 3A 2 yo 4- A 3 yo- 

These results can be written symbolically as 

Vi = (1 4- &)yo, 2/2 - (1 + &) 2 Vo, 2/3 8=8 (1 4- A) 3 yo 

in which (1 4~ &) k is an operator on y 0 with the exponent on the A indicat- 
ing the order of the difference. The difference operator A is analogous to 
the differential operator D introduced in Chap. 1. 

We easily establish by induction that 

y k - (1 + A) k y 0t *- 1,2,..., (6-3) 

or, in the expanded form, 

k(k - 1) , k(k - l)(fc - 2) . 

Vk - Vo + k Aj/o -4 — A 2 j/ 0 4 — A 3 y 0 4 . (6-4) 

Formula (6*4) enables us to represent every value yk in terms of y 0 and 

the forward differences A y 0 , A 2 y 0> 

We can derive a similar formula by starting with the values of the yB 
at the end of Table 1 and forming the backward differences defined as 
follows: The first backward differences Vyi are 

Vyi * Vz - 1 * (6-5) 

The second backward differences V 2 y x are defined by 

V 2 y, = Vy x - ( 6 - 6 ) 

and in general, the kth backward differences V k y % are 

vV = (6-7) 

1 A differences table in a specific numerical example appears in the Example of the 
next section. 



SEC, 81 INTERPOLATION AND EMPIRICAL FORMULAS * 693 

A table of backward differences is indicated in Table 2, where the dif- 
ferences V k yi with a fixed subscript i lie along the diagonals slanting up, 
as shown by arrows. 

Table 2 



Now, from (6-5) to (6-7) we deduce that 

Vn ~ ^ Vn ^ Vn—l ~ Vn tyri-l H" Vn—2 

V z Vn = v 2 y n - v 2 y n - 1 * Vn - 32/ n ~i + 3t/„-2 - y n ~3 
and in general k 

V k y n = V fc_1 J/r. ~ = 23 ( _1 ) r ( ) Vn—ri 

r«*0 \r/ 

k(fc - l)(fc - 2) ... (fc - r + 1) 


where 


0 


(6-8) 

(6-9) 


is the binomial coefficient of x r in the expansion of (1 + x) k , 

By using (6-8) successively in the definitions of backward differences 
we find 

2/n~ 1 * Vn ~ Vyn s (1 “ V)Vn, 

y n -~2 " Vn 2 Vy n + V 2 2 / n “ (1 “ y)*Vni 


and, in general, ^ - (1 - V)*y«, (6-10) 

where V is the backward-difference operator. The formula (6-10) when 




NUMERICAL ANALYSIS 


[chap. 9 


expandedreads 

Vn-k “ Vn - kvy n + 


k(k - 1 ) 
2! 


V 2 ?/„ 


*(* - l)(Jfc - 2) 

3! 


( 6 - 11 ) 


It shows that any value of y in Table 2 can be expressed in terms of y n 
and backward differences V h y n . 

We shall use formulas (6-4) and (6-11) to derive certain interpolation 
formulas and to deduce some formulas for numerical integration. 


PROBLEMS 


1. Compute the forward and backward differences for the following set of data; 


X 

l 

2 

3 

4 

5 

6 

7 

8 

y 

2.106 

2.808 

3.614 

4.604 

5.857 

7.451 

9.467 

11.985 


2. Write expressions for the yu, k ** 1, 2, . . ., in Prob. 1 by using (6-4) and (6-11). 


7. Polynomial Representation of Data. Unless a statement to the con- 
trary is made, we shall suppose henceforth that the values x x in- a given 
set of data (x,-,^*), where % ** 0, 1, 2, . . n, are equally spaced. If the 
spacing interval is h, then 

X\ = Xq “f* h } X 2 — Xq -j- 2 h, . . . , X n s* Xq -f~ nh. 

We pose the problem of representing the data by some formula y = f(x), 
which for x = xq + kh yields yk — f(x 0 + kh). We shall frequently write 
fk for y k . 

We observed in the preceding section that whenever the rth differences 
of the ys are constant, then all differences of order higher than r vanish. 
In this event formula (6-4) yields 


Vk « 2/o + 




where the binomial coefficients 



are defined by 


(7-1) 


0 k(k - l)(k - 2) r + 1) 

r! 


(7-2) 


Since the x x are spaced h units apart, 

x k =* Xq + kh, k « 1, 2, . . n, 
x k — Xq 

so that k * — — (7-3) 

h 

Now the expression (7-2) is a polynomial of degree r in k. Therefore, on 



f 

8EC. 7] INTERPOLATION AND EMPIRICAL FORMULAS 695 

substituting in (7-1) for k from (7-3) we obtain a polynomial of degree r 
in Xk . When like powers of Xk are collected, (7-1) takes the form 

Vk *= <*0 + a\Xk + a 2 xl 4 b d r xf k . (7-4) 

Accordingly, the polynomial in x, 

y(x) « oo + aix + a^ 2 4 f 0 ^, (7-5) 

assumes the values 2 /* when we set x « x*. Thus, when the rth differences 
of the pk are- constant and the Xk are equally spaced, the polynomial (7-5) 
represents these data exactly. 

It is easy to prove a converse to the effect that the rth differences of 
the polynomial (7-5) are constant. It would suffice to show that the 
first difference Ay(x) = y(x + h) — y(x) formed with the aid of (7-5) 
is a polynomial of degree r — 1, for if differencing a polynomial once re- 
duces its degree by 1, r successive differencings would yield a polynomial 
of degree 0, that is, a constant. 1 

When rth differences in a given set of data are not constant but differ 
from one another by negligible amounts, the polynomial (7-5) represents 
the data approximately. 

Example: The set of data and the forward differences tabulated below suggest that 
these data can be represented by a cubic polynomial y — ao -f aix + a*x 2 -f- a&x* if 
two-decimal accuracy is sufficient. 



1 We leave it to the reader to show that Ay(x) is, indeed, a polynomial of degree r — 1. 
The result is analogous to the theorem that the derivative of a polynomial of degree r 
is a polynomial of degree r — 1. The expression Ay = y{x -f h) — y{x) save for the 
factor 1/A is the difference quotient used in defining the derivative. 






896 NUMERICAL ANALYSIS [CHAP. 9 

The coefficients a in this polynomial can be determined with the aid of formula (7-4) 
by using (7-1) with r » 3 and by taking 

l/o m 2.105, Aj/o « 0.703, A Vo ” 0.103, A 8 j/o m 0.081. 

Since such calculations present no interest, we do not give them here. It is more sensible 
to determine the o» by the method of least squares of Sec. 11. 

PROBLEMS 

1. Given the table: 


X 

19 1 

20 

21 

22 

23 

24 

25 

y 

81.00 j 

90.25 

100.00 

110.25 

121.00 

132.25 

144.00 


Compute second forward differences, and represent the data by y ** ao -f aix 4* o^x 2 . 
Determine ao, «i, 02 so that the polynomial passes through (a) the first three points, 
(b) the last three points. 

2. Discuss the calculation of the 1 /* in Prob. I from (6*4) and (6-11). 

8. Newton’s Interpolation Formulas. When the data {x i ,y l ) } where 
i 0, 1, 2, . . ft, are presented in tabular form, an infinite numl>er of 
analytic relations y =*= f(x) can be devised such that iji = /(x») either 
exactly or approximately. Once a suitable form of /(. r) is determined, the 
formula y = f(x) can be used to calculate the ordinates y for xs not ap-* 
pearing in the table. That is, the formula can be used for interpolation 
or extrapolation. 

The simplest of such formulas is a linear relationship based on the 
assumption that the values of y in the interval (x t -,x t+l ) can be represented 

by 

V = Vi + — - — — (x - Xi). (8-1) 

Formula (8-1) is precisely that used in estimating the values of such tabu- 
lated functions as logarithms by the process of “interpolation by pro- 
portional parts.” 

More accurate interpolation formulas are based on the assumption that 
the desired value of y can be computed from a polynomial 

y * a 0 + a x x + a 2 x 2 4 f a m x m (8-2) 

in which m + 1 coefficients a»* arc so chosen that m + 1 pairs of tabulated 
values (Xi,y x ) satisfy (8-2) exactly. 1 

In the preceding section we saw that when the data are represented by a 

1 These m + 1 pairs may include the entire set of given values (x*,y»), or they may 
be a subset so chosen that \x - Xi | is as small as possible. 



t 


SEC. 8] INTERPOLATION AND EMPIRICAL FORMULAS 607 


polynomial of degree m, then all forward differences of order higher than 
m vanish. Accordingly, formula (6-4) yields 


Vk 


m - 1) _ 

Vo + k A^/o 4 r: A^yo + * 

/ 1 


• + 


k(k — 1) . ♦ . (ft — m + 1) 
m\ 


& m yo 

(M) 


and, since the x, are equally spaced, x k — x 0 + kh, so that 




On inserting this value of k in (8-3) we get 


x k - x o (x k 

V* = 2/o H ; Aj/o 4 


TqXt* - r 0 - h) 

2\h 2 


A 2 2/o + ■ 


+ 


(xk ~ r 0 )(Xk - x» - ft) ■ . ■ Or* - Jo - mh -f h) 
_____ 




(8-4) 


This relation is satisfied by m + 1 pairs of the tabulated values. If 
we assume that the value of y corresponding to an arbitrary x can be 
obtained from (8-4) by replacing t* by x, we get the formula 


vM = ?/o + 



Aj/o + 


(j — r 0 )(x - x 0 — ft) 

Jih 2 


A 2 y 0 H 


+ 


(t - T 0 )(.r — x 0 — h) ... (x - T 0 - mJi + A) 

»7w” 


A M J/o 


(8-5) 


known as Newton' s forward-d iff ere nee interpolation formula. This formula 
can, of course, he used for either interpolation or extrapolation. 

By replacing (jc — Jto)/h by a dimensionless variable A" which represents 
the distance of x from x 0 in units of h, we get from (8-5) 


Vx 


Vo + X Ay 0 + 


X(X - 1) 

2 ! 


A 2 ?/o 4 


, X(X- 1) ... (X-m+1) 

+ : A m y 0 , 

ml 


(8-6) 


where X = (.r — or 0 )//i and = y(x 0 + hX) — j/(x). 

A similar calculation based on the use of (6-11) yields Newton's backward- 
difference interpolation formula 


X (X + 1) 

Vn+x Vn + X Vy n H ~ V 2 y n + * 


, X(X + 1) . . . (X + m - 1) _ 

+ : V> n (8-7) 

m! 



NUMERICAL ANALYSIS 


[chap. 9 


where 



so that x = x n + XX, 


and Vn+x = y(x n + hX ) = y(x). 

When the data cannot be represented by a polynomial, the right-hand 
members of (8-6) and (8-7) are infinite series involving differences of all 
orders. 

Formulas (8-6) and (8-7) can be used to compute derivatives of tabu- 
lated functions. Thus, on differentiating successively (8-5) with respect 
to x and setting x — x 0 in the result, we get 




i / , , n l 5 . \ 

y"{x o) - — 2 ^A 2 !/o - A 3 y 0 + — A 4 j/o “ - A 5 2 /o H J 

V"\xo) = ( A *»o ~ ^ A 4 t/ 0 + ^ A 5 t / 0 ) 


( 8 - 8 ) 


V W (x 0 ) = — (A 4 ?/ 0 - 2A 4 j/o 4 )• 

Formulas (8-8) should be used with caution because even when y = f(x) 
is well represented by the polynomial P(x) f the derivatives of f(x) may 
differ significantly from those of P(x). 

Example : Using the data given in the Example in Sec. 7, determine an approximate 
value for the y corresponding to x * 2.2. 

First, let y be determined by using only the two neighboring observed values (hence, 
m » 1). Then, xq * 2, yo *■ 2.808, Ayo ** 0.806, and X «* (2.2 — 2)/l » 0.2. Hence, 

y « 2.808 -f 0.2(0.806) * 2.969, 

which has been reduced to three decimal places because the observed data are not given 
more accurately. This is simply a straight-line interpolation by proportional parte. 

If the three nearest values are chosen, m » 2, ro ~ 1, yo 88 2.106, Ayo 0.703, 
A*y 0 » 0.103, and X « 2.2 - 1 - 1.2. Then, 


y - 2.105 + 1.2(0.703) + ( ( 0 .103) - 2.961, 
correct to three decimal places. 

If the four nearest values are chosen, m «* 3, xq « 1, j/o • 2.105, Aye •» 0.703, 
A*yo ** 0.103, A*yo ■» 0.081, and X ** 1.2. Therefore, 


v - 2.105 + 1.2(0.703) + — - (0.103) + (1 ' 2 ^ 0 ' 2)( °' 8) (0.081) - 2.958, 
2 6 


correct to three decimal places. 



f 


SEC. 9] INTERPOLATION AND EMPIRICAL FORMULAS 699 

PROBLEMS 


1. Compute with the aid of formulas (8*6) and (8-7) the approximate values of y 
corresponding to x «* 6.6 from the data of the Example in Sec. 7. Use two and three 
neighboring values. 

2. Extrapolate the value of y for x » 8.2 from the data in the Example of Sec. 7 with 
the aid of (a) formula (8-6), (b) formula (8-7). Use m ** 2. 

3. Compute j/'(l) and y”( 1) from the data of the Example of Sec. 7 with the aid of 
( 8 - 8 ). 


9. Lagrange’s Interpolation Formula. The interpolation formulas de- 
veloped in the preceding section apply only when the given set of x x is 
an arithmetic progression. If this is not the case, some other type of 
formula must be applied. 

As in Sec. 8, select the m + 1 pairs of observed values for which \x — x t \ 
is as small as possible, and denote them by (x x ,y t ) where i « 0, 1, 2, . . . , ra. 
Let the rath-degree polynomials Pk(x) y where k ~ 0, 1, 2, . . ., ra, be de- 


fined by 


(x - x 0 )(x - Xi) ... (x- x m ) 
x - X k 


IK*- *»)• 


Then, the coefficients Ak of the equation 


t r*k 


y = 


E A k Pk(x) 


0 


can be determined so that this equation is satisfied by each of the ra - f 1 
pairs of observed values (r„y t ). For if £ = Xk, then 


since P*(r t ) 


Ak 


Vk 

Pk(x k ) 


0 if i 5 ^ k. Therefore, 


A VkPk(x) 

v = y. 

to Pk(Xk) 


(9-2) 


is the equation of the rath-degree polynomial wdiich passes through the 
ra + 1 points whose coordinates are (x xy y x ). If x is chosen as any value 
in the range of the x ly (9-2) determines an approximate value for the 
corresponding y. 

Equation (9-2) is known as Ixigrange’s interpolation formula , Ob- 
viously, it can be applied when the x t are in arithmetic progression but 
(8-5) is preferable in that it requires less tedious calculation. Since only 
one rath-degree polynomial can be passed through m + 1 distinct points, 
it follows that (8-5), or its equivalent (8-6), and (9-2) are merely different 
forms of the same equation and will furnish the same value for y. 



700 

Example: Using the data 


NUMERICAL ANALYSIS 


{chap. 9 


V 

10 

15 

22.5 

33.75 

50.625 

75.987 

p 

0.300 

0.675 

1.519 

3.417 

7.689 

17.300 


apply Lagrange's formula to find the value of p corresponding to v — 21. 

II the two neighboring pairs of observed values are chosen so that m ** 1, 


V 


« 0.675 


21 - 22.5 
15 - 22.5 


+ 1.519 


21 - 15 
22.5 - 15 


1.350, 


correct to three decimal places. 

If the three nearest values are chosen so that m ■* 2, 


rto {21 - 15)(21 - 22.5) , (21 - 10)(21 - 22.5) 

V * u.o r 0.075 

V (10 - X 5) ( 10 - 22.5) (15 - 10)(!5 - 22.5) 


correct to three decitn&l places. 


+ 1.519 


(21 - 10)(21 - 15) 
(22J- 10K22.5 - 15) 


1.323, 


PROBLEMS 

1. Using the data of the Example in Sec. 9, find an approximate value for p when 
v m 30. Use rn » 1 and m » 2. 

2. Use m ■ 1, 2, and 3 in formula (8-6) to find an approximate value of 0 when 
t « 2.3, given 


t 

0 

1 i 

2 

3 | 

4 

5 

6 

7 

8 

& 

60.00 

51.66 

44.46 

38.28 

32.94 

28.32 

24.42 

21.06 

18.06 


3 . Given the data 


X 

0.16 

0.4 

1.0 

2.5 

6.25 

15.625 

V 

2 

2.210 

2.421 

2.661 

2.929 

3.222 


find an approximate value of y corresponding to x « 2. Use formula (9-2) with m » 1 
and mm 2. 

4 . Given the data 


C 

19 

20 

21 

22 

23 

24 

25 

H 

81.00 

90.25 

100.00 

110.25 

121.00 

132.25 

144.00 


find an approximate value of H when C » 21.6. Use formulas (8-6) and (8-7) with 
m «* 1, 2, and 3, 



SEC. 101 INTERPOLATION AND EMPIRICAL FORMULAS 701 

10, Empirical Formulas. A given set of discrete data can be represented 
analytically in infinitely many ways. Such analytic representations are 
called empirical formulas , and the choice of the functional form for an 
empirical formula ordinarily depends on the use to be made of the formula. 
Thus, if a given set of data is to be represented by a function f(x) which 
enters in the differential equation 

Liu) -/(*), 

the form of fix) may well depend on the ease with which this equation can 
be solved. For some types of differential operators L it may be wise to 
take f{x) as an algebraic polynomial, in others as an exponential, and so 
on. Because of the commonness of algebraic and trigonometric poly- 
nomials in applications, we confine our discussion of empirical formulas 
primarily to these two types. 

The first step usually taken by an experimenter in appraising a set of 
observed values is to plot them on some coordinate paper and draw 

a curve through the plotted points. If the points when plotted 

on a rectangular coordinate paper, lie approximately on a straight line, 
he assumes that the equation y — mx + b represents the relationship. 
To determine the constants m and b, the slope and the y intercept may be 
read off the graph or they may be calculated by solving two linear equations 
for m and b got by substituting the coordinates of two judiciously chosen 
points on y ~ mx 4* b. 

If the plot of points on a logarithmic coordinate paper indicates that they 
lie on a straight line, the desired relationship has the form 

y = ax m } 

for on taking logarithms, we get 

log y * log a + m log x y 

and if coordinate axes X, Y are marked so that log y « Y and log x = X, 
we get a linear equation 

Y = log a + mX. 

Again the constants a and m can be either read off the graph or computed 
by solving a pair of linear equations for m and log a. 

Similarly, the data can be represented by an exponential function 

y « alO m * 

if the values (x^yi) when plotted on a semiiogarithmic paper fall on a 
straight line, for on taking logarithms to the base 10, we get 

log y » log a + mx , 

which is linear in log y and x . 



702 


NUMERICAL ANALYSIS 


[CHAP. 9 


When none of these simple functional relationships fits the data, one 
may determine, with the aid of Sec. 7, if the data can be fitted by a poly- 
nomial. It should be stressed, however, that ordinarily the choice of an 
empirical formula is governed by whatever uses are to be made of it. Once 
a formula is chosen, the parameters entering in it (such as the coefficients 
in the polynomial representation) can be determined by imposing some 
criterion for the goodness of fit of the data by the chosen function. The 
method of least squares, presented in the next section, provides one of the 
most commonly used of such criteria. 


PROBLEMS 

1. Plot the following data on a rectangular, logarithmic, or semilogarithmic paper to 
determine the approximate functional relationships between y and x. 


X 

3 

4 

5 

6 

7 

8 

9 

10 

11 

12 

y 

X 

5 

! 1 

1 5 

j 

.6 

1 2 

6 

M 

6.4 

3 

7 

4 

7.5 

5 

8.2 

6 

8.6 

7 

9 

8 

9.5 

j 9 

y 

2.5 

3.5 J 

4.3 

5 

5.6 

6.2 

6.6 

7.1 

7.5 


X 

l 

2 

3 

4 

5 

6 

7 

8 

y 

0.5 

0.8 

1.2 

1.9 

3 1 

4.8 

7.5 

11.9 


2 . Verify that the data in Probs. 2, 3, and 4 in Sec. 9 may be approximated by the 
following types of functions: 0 — a 10"**, y « ax m , H « ao -f -f C 2 , respectively. 
Determine the parameters graphically or analytically. 

11. The Method of Least Squares. We saw in Sec. 7 that the m + 1 
coefficients in the polynomial 

2/ ~ ao + aix H h a m x m (11-1) 

can always be determined so that a given set of m + 1 points ( 
where the xs are unequal, lies on the curve (1 1-1). When the are equally 
spaced, the desired polynomial is determined by the formula (8-5) and, 
in the more general case, by (9-2). 

When the number of points is large, the degree m of the polynomial 
(11-1) is high, and an attempt to represent the data exactly by (11-1) 
not only is laborious but may be foolish, for the experimental data in- 
variably contain observational errors and it may be more sensible to rep- 
resent the data approximately by some function y * f(x) which contains 
a few unknown parameters. These parameters can then be determined 
so that the curve y « f(x) fits the data in “the best possible way.” The 



INTERPOLATION AND EMPIRICAL FORMULAS 


703 

are, of course, 


sec. 11] 

criteria as to what constitutes “the best possible way 1 
arbitrary. 

For example, we may attempt to 
fit the set of plotted points in Fig. 12 
by the straight line 

y « a x + a& 

and choose the parameters a\ and a 2 
so that the sum of the squares of 
the vertical deviations of the plotted 
points from this line is as small as 
possible. 

More generally, if we choose to 
represent a set of data (x v ,y x ), where 

t =^* 1, 2, . . ., n, by some relationship y — f(x ), containing r unknown 

parameters a t , a 2 , . .., a r , and form the deviations (or the residuals , as 

they are also called) ... m 

3 (H-2) 



the sum of the squares of the deviations 

vt = T, m*i) - yH 2 

t-i 


(11-3) 


is clearly a function of a% f a 2 , . . a r . We can then determine the as so 
that S is a minimum. 

Now, if *S(aj,a 2 f . . ,a r ) is a minimum, then at the point in question 


as 

as 

— = 0, 

— 

a«i 

da 2 


dS 

da r 


(11-4) 


The set of r equations (11-4), called normal equations , serves to determine 
the r unknown as in y ~ f{x). This particular criterion of the “best fit” 
of data is known as the principle of least squares, and the method of de- 
termining the unknown parameters with its aid is called the method of least 
squares . It was introduced and fully developed by Gauss 1 when he was 
a youth of seventeen! 

We indicate the construction of the normal equations first by supposing 
that y = f(x) is a linear function 

y » a x + a&. (11-6) 

1 The criterion of least squares plays a fundamental role in the approximation of a 
suitably restricted function f(x) by a linear combination of orthogonal functions. As is 
shown in Chap. 2, Sec, 23, the partial sums of Fourier series give the best fit in the sense 
of least squares. It should be noted, however, that the polynomials giving the best 
fit to f(x) in the sense of least squares, in general, are not the partial sums of Maclaurin’s 
or Taylor’s series for/(x). 



NUMERICAL ANALYSIS 


(chap. 9 


704 

The residuals (11-2) for (11-6) are 

Vi * (ai + a a xi) - yu 

so that S =* 2 v < 

i- 1 

= (ai + 02X1 — y{f + (ai + 02X2 — 3/2)* 

+ • • • + (oi + «2X n — y„) 2 . 

On differentiating S with respect to a x and a 2 , we deduce two equations: 


— = 2 (a! + 02X1 - Vi ) + 2 (aj + a 2 x 2 - 2/2) 

ckii 

-1 h2(o! + a 2 x„ - jf„) = 0, 

SS 

— = 2x!(a! + « 2 X 1 - j/0 + 2x 2 (a 1 + < 12 X 2 - y 2 ) 
da-t 

-j h 2x„(o, + a 2 x„ - 2 /„) = 0. 

If we divide out the factor 2 and collect the coefficients of a x and o 2 , we get 

nai + ( E x i ) «2 = £ Vi, 

(n-6) 

( E X,') flj + { E *i) <*2 = E *#.'• 

\t=*i / / i»»i 

These equations can be easily solved for ai and a 2 . 

Exavi-ple 1. We illustrate the use of Eqs. (11-6) by calculating the coefficients in 
y » a\ *f a& to fit the following data: 


X 

i 

2 

3 | 

4 

y 

1.7 

1.8 

2.3 

3.2 


In this case n «* 4, and since 

4 

Ex. - 1 +2 + 3 + 4 - 10, 

1-1 

4 

Es«“ 1.7 + 1.8 + 2.3 + 3.2 -9, 

»«*1 

4 

E A - 1 + 4 + 9 + 16 - 30, 

1-1 

4 

E “ 1-7 + 2(1.8) + 3(2.3) + 4(3.2) - 25, 

tml 



705 


BBC. 11] INTERPOLATION AND EMPIRICAL FORMULAS 

the system (11-6) reads 

4ai -f lOaj « 9, 

10a i -f 30a 2 « 25. 


Solving for ai and a% we get ai « 1, as « l A> so that the desired straight line fitting the 
data in the sense of least squares is y « 1 4* Ax. 


We suppose next that y = f(x) is a polynomial 

y ~ ai + a 2 x + a 3 x 2 H f- a r x r ~ l 

= X) 


The residuals Vi this time are 


Vi = 2 <*,*{ 1 - Vi. 


(11-7) 


(11-8) 


Since 


s = £)»?, 

t-1 


Eqs. (11-4) can be written as 


From (1 1-8), 


as A 

” 2 23 r ““ 0i k «* 1, 2, . . , , r. 

aajk ,»i oak 


— = r*- 1 
dajk 


(11-9) 


so that, on dividing out the factor 2, we can write the normal equations 
(11-9) as 

n 

£ e,*? -1 = 0. (11-10) 

l-l 

The substitution from (11-8) in (11-10) yields 

- Vi) rf" 1 = o, 

•«*i i / 

and on collecting the coefficients of the a ; -, we get a set of r linear equations 

£ (£ 4 +k ~ 2 } ay = £ *f'V. fc = 1, 2, . . r, (11-11) 

J— 1 V—l / i-l 

for A| f ®2j • * *y ftr* 

We illustrate the use of these equations by two examples. 



705 NUMERICAL ANALYSIS [CHAP, 9 

Example 2, Let the data in Example 1 be fitted by y * ai 4* <*& 4* a& l < Then 

Vi m ai 4* flax* 4~ - Vi 


and 

The normal equations 


Bv t dVi dv » 

»* 1 «* Xi, — «■* xf. 

da\ dot das 

£><— -0, jt-1, 2, 3, 

S3 do* 


53 (°i + °**» + - j/») *1 *0, 


53 (oi -f W *f osaf - Vi)x t - 0, 

*— l 


53 (°i + <*&* 4- smf - P.)4 - o. 

4-1 

If the coefficients of the a, are collected and the normal equations put in the form 
(11-11), one obtains the three equations 



** + (,?,*■) 

at + ^53 *<) 03 

4 

~ Ei+ 

»-l 


(S*‘)" + (S 1 ») 

as 4- (l3 x ?) as 

4 

* 13 X W, 


(S'*)*' + (S-0 

«2 4* (53 art) «3 

- E A vi- 

i-i 

X> 
<- 1 

«14-24-34-4*10, 

53 A * 1 4- 

»-i 

4+9 + 16 

4 

53 *<3/* 

• 1.7 4*3.6 4“ 6.9 4*12 

.8 ** 25, etc. 



Now, 


t—i 

The equations become 

4ai -f IO 02 -I- 30aa - 9, 

10ai + 3002 -f 100a 3 » 25, 

30oi + 100a* 4- 354a 8 * 80.8; 

and the solutions are a\ ** 2, 02 *■ —0.5, oj » 0.2. 

Example 8. Let us apply the method of least squares to fit the data 


X 

1 

2 

3 

4 

5 

6 

7 

8 

V 

2.105 

2.808 

3.614 

4.604 

5.857 

7.451 

9.467 

11.985 


by the polynomial y - ai + 4- 0 * 2 * 4- «43*. 




SEC. 11 } INTERPOLATION AND EMPIRICAL FORMULAS 707 

In this case n «* 8 and Eqa. (11*11) yield four normal equations obtained by setting 
k — 1, 2, 3, 4. They are 

8o! + ( E *<) “s + ( E *? ) a > + ( E *?)«*“ JL Vi, 

(Ex,) oi + (Z 4 ) or + (Ex?) O, + (gx?) «« - 

(Ex?) Ol + (Ex?) o, H- (E x{) O, -1- (Ez?) 04 - Ex?«i, 

(Ex?) 01 + (Ex?) 02 + (Ex?) oj + (Ex?) 04 - E4w- 

From the form of the coefficients of the a*, it is seen that it is convenient to make a 
table of the powers of the x % and to form the sums and y % before attempting to 
write down the equations in explicit form. 


MS 

x? 

3 

Xi 

4 

Xi 



1 

1 

1 

1 

1 

1 

2 

4 

8 

16 

32 

64 

3 

9 

27 

81 

243 

729 

4 

16 

64 

256 

1,024 

4,096 

5 

25 

125 

625 

3,125 

15,625 

6 

36 

216 

1,296 

7,776 

46,656 

7 

49 

343 

2,401 

16,807 

117,649 

8 

64 

512 

4,096 

32,768 

262,144 

Sxi 36 

204 

1,296 

8,772 

61,776 

446,964 


fl 

Vi 

x<Vi 

Ay, 

r— 

4vi 

1 i 

2.105 

mm 

2.105 

2.105 

2 

2.808 


11.232 

22.464 

3 

3.614 

10.842 

32.526 

97.578 

4 


18.416 

73.664 

294.656 

5 

5.857 

29.285 

146.425 

732.125 

6 

7.451 

44.706 

268.236 

1,609.416 

7 

9.467 

66.269 

463.883 

3,247.181 

8 

11,985 

95.880 

767.040 

6,136.320 

Zx\y x 

47,891 


1,765.111 

12,141.845 














70S NUMERICAL ANALYSIS [CHAP, 9 

When the values given in the tables are inserted, the normal equations become 
80 ! + 360* + 204oa + l ( 296<n - 47.891, 

36di + 204oj + l,296o 3 + 8,772en - 273.119, 

204aj + 1, 29602 + 8,772 a, + 61,776 a« - 1,766.111, 
l,296oi + 8,772os + 61,776o s + 446, 964a* - 12,141.846. 

The solutions are 


<H - 1.426, as - 0.693, o 8 * -0.028, 04 - 0.013. 

Therefore, the equation, as determined by the method of least squares, is 
y « 1.426 + 0.693x - 0.028s 1 2 + 0.01 3** 

The normal equations ( 11 - 11 ), corresponding to the polynomial repre- 
sentation of data, are linear in the coefficients a % . They need not be linear 
in the unknown parameters if the function y = f(x) is not a polynomial 
in x. In this event the solution of the system (11-4) may prove difficult, 
and one may be obliged to seek an approximate solution by replacing the 
exact residuals ( 11 - 2 ) by approximate residuals which are linear in the 
unknowns. This is accomplished by expanding y — f(x), treated as a 
function of a X) a 2 , . a r , in Taylor’s series in terms of a x ~ d x ss Aa t , 
where the are approximate values of the a The values of a % may be 
obtained by graphical means or by solving any r of the equations y t — }{x x ). 
The expansion gives 

y « f(x , a u . . a r ) » f(x, d x + A a u . . d r + Aa r ) 


where 


' df 

1 /(^j #1> • * • y &r) + 2 ~ A a k 
k -1 V a k 

1 ^ d 2 / 

+ ~ 2^ Aa ) Aa * H — 7 


2 !/,*«! ddjddk 


( 11 - 12 ) 


df df 

— s — 
ddfc dajc 


d 2 f 


d 2 f 


ddj d&)c ddj dak 


etc. 




Assuming that the d t are chosen so that the A a, are small, the terms of 
degree higher than the first can be neglected and ( 11 - 12 ) becomes 


y ” f(x,a lf . . .,Or) + 


E — 

Jfc«i ddk 


Aa k . 


The n observation equations are then replaced by the n approximate 
, equations 


df 



SAC* II] INTERPOXtATION AND EMPIRICAL FORMULAS 700 

If (11-13) is used, the residuals Vi will be linear in the Aa*, and hence the 
resulting conditions, which become 

&S 

— — - = 0, k - 1, 2, . . r, (11-14) 

d(Aa k ) 

also will be linear in the Aa *. Equations (11-14) are called the normal 
equations in this case. 

We illustrate the use of Eqs. (11-14) in Example 4. 

Example 4. We seek to determine the constants k and a in the formula 0 * ka* 
chosen to represent the following data: 


t 

l 

2 

3 

4 

e 

51.60 

44.46 

38.28 

32.94 


The determination of A: and a in this problem can be reduced to the solution of two 
linear equations, for if we write B «* ka * in the form 

e - mo* 

then on taking logarithms to the base 10, we get 

log B ® log k 4* bt. 

Setting log $ «* y and log k ** K, we get 

y ** K + bt (11-15) 

which is linear in K and b. These constants can be determined by the procedure de- 
scribed above, which leads to the solution of a pair of linear equations. 1 

To illustrate the use of formulas (11-12) to (11-14), we follow a more laborious route 
which gives an approximation to the original equation. 

When the values recorded in the table are plotted on semilogarithmic paper, it is 
found that A; — 60 and a ** 10~ 0 068 «* 0.86, approximately. This suggests using k$ » 60 
and oo * 0.9 as the first approximations. The first two terms of the expansion in 
Taylor’s series in terms of Ak =* A; — 60 and Aa « a — 0.9 are 

B - 60(0.9) 4 + o Afc + Aa 

\dK/ a—0 9 \oa/a«-0.9 

- 60(0.9)* 4- (0.9)* AA: + 60*(0.9)*~ l Aa. 

If the values (UA) are substituted in this equation, four equations result, namely, 

0i - 60(0.9)** + (0.9)** Ak + 60<.(0.9)**“ 1 Aa, * - 1, 2, 3, 4. 

The problem of obtaining from these four equations the values of AA; and Aa, which 
furnish the desired values of is precisely the same as in the case in which the original 
equation is linear in its constants. The residual equations are 

Vi - (0.9)** Ak + 60tj(0.9)**~ 1 Aa + 60(0.9)** - 9i, i - 1, 2, 3, 4. 

1 However, the approximation obtained by this meanfl does not give an approximation 
to the original equation in the sense of least squares. 



710 

Therefore 


NUMERICAL ANALYSIS 


$ » E - Z 1 ( 0 . 9 ) '• A k + 60 « 0 . 9 )'<~ l Aa + 60 ( 0 . 9 )*< - #< J* 

»—l t— 1 


[chap. 9 


and the normal equations 


become 


and 


AS 

d(A k) 


- 0 and 


AS 

d(Aa) 


• 0 


2 £ [0.9‘* A* -f WCO.Q)^- 1 Aa + 60(0.9)** - - 0 

i-X 


2 X) [0.9 f * A/e + 60^(0 S) 1 *- 1 A a + 60(0.9)^ - ftjeO^O^) 1 - 1 « 0. 
*-x 


When these equations are written in the form 

p A/e + q Aa - r, 

with all common factors divided out, they are 


and 


E ( 0 . 9 )"< AJb + 60 E {.(O.O) 11 *- 1 Ao - E #i( 0 . 8 )** - 00 E ( 0 . 9 )*‘< 

**•1 t««l i»l 

4 4 4 4 

E t<( 0 . 9)**<— 1 A* + 60 E f?( 0 . 9 ) 5i -~ ! AO - E #rf>( 0 . 9 ) ,,_l - 60 E ‘.(O.O) 1 **- 1 . 

«*»1 *»-l *X»1 *—l 


As in Example 3, the coefficients are computed most conveniently by the use of a table. 


mm 

a 

2 

3 

4 

Totals 

(0.9) '< 

0.9 

0.81 

0.729 

0.6561 


(0.9)*“ 


0.6561 

0 531441 

0.43046721 

2.42800821 

tj(0.9)*“ -1 

1 

1.458 

3.77147 

1.9131876 

6.0426576 

tf(0.9)’**-‘ 

1 

3.24 

5.9049 

8.503056 

18.647956 

(«.)(0.9)“ 

46.494 

36.0126 

27.90612 

21.611934 

132.024654 

mm)*- 1 

51.66 

80.028 

93.0204 

96.05304 

320.76144 


Substituting the values of the sums from the table gives 


and 


2.42800821 A k + 362.659456 Aa - 132.024654 - 145.6804926 
6.0426S76 A k + 1,118.87736 Aa - 320.76144 - 362.559466. 


























BBC. 12] INTERPOLATION AND EMPIRICAL FORMULAS 711 

Reducing all the numbers to four decimal places gives the following equations to solve 
for Ak and Aa: 

2.4280 Ak + 362.5595 A a - -13.6558, 

6.0427 Ak + 1,118.8774 Aa - -41.7980. 

The solutions are 

Ak m —0.238 and Aa « -0.036. 

Hence, the required equation is 

6 « 59.762(0.864)*. 

PROBLEMS 

1. Apply the method of least squares to find the constants in y *« ai 4* <* 2 # 4* ajjx 2 
to fit the data 


X 

l 

2 

3 

4 

5 

6 

y 

3.13 

3 76 

6.94 

12.62 

20.86 

31.53 


2. Determine by the method of least squares the constants a and n in p « av n to 
fit the following data by writing the equation in the form 

log p » n log v 4 log a. 


V 

10 

15 

22.5 

33.7 

1 50.6 

75.9 

V 

0.300 

0 675 

1.519 

3.417 

7.689 

17.300 


Hint: Set log p « y y log v ~ x , and determine the constants in the resulting linear 
equation. 

8. Compare the result of Example 4 with the calculation of the constants in (11-15). 

12. Harmonic Analysis. The problem of representing a suitable periodic 
function in a trigonometric series was considered in some detail in Chap. 
2. In this section we give a brief discussion of the problem of fitting a 
finite trigonometric sum to a set of observed values Let the set 

of observed values 

ip^OiV o)> (*£l>2/l)> • • •> (X2n—li2/2n— 1)> (% 2n)V2n)t 

be such that the values of y start repeating with y 2n (that is, y 2n « yo , 
y 2n +i = Vu etc.). It will be assumed that the x t are equally spaced, that 
x 0 = 0, and that x 2n = 2*. [If x 0 ^ 0 and the period is c instead of 
2w, the variable can be changed by setting 

2w 

==: (x* Xo)* 

c 

The discussion would then be carried through for and j/,- in place of the 



T12 NUMERICAL ANALYSIS 

X{ and Vi used below.] Under these assumptions 


[chap. 0 


Xi ** 



VK 

n 


The trigonometric polynomial 

« n — 1 

y «* A 0 + A* cos kx + 2 s* n kx (12-1) 

Jfc-1 Jk-4 

contains the 2n unknown constants 


A(); ^1) A-2i • • •> A n , Bl) &2, • • •> -Bn-1, 

which can be determined so that (12-1) will pass through the 2 n given 
points (x*,y,) by solving the 2n simultaneous equations 

n n~l 

yi * A 0 + £ A* 008 kx < + Yj B k sin kx if i * 0, 1, 2, . . 2n - 1. 

1 *«*1 


Since $,* «* tir/n, these equations become 

" ikr ihv 

V% » A 0 + Za cos b 2-, sin — » 

k-~i n jk„i n 


i - 0, 1, 2, 2n - 1. (12-2) 


Hie solution of Eqs. (12-2) is much simplified by means of a scheme 
somewhat similar to that used in determining the Fourier coefficients, 
Multiplying both sides of each equation by the coefficient of A 0 (that is, 
by unity) and adding the results give 


* Z 2 ^ 1 *r\ V? Z 2 ^ 1 ik *\ 

£ Vi *= 2ftA 0 + ]C ( cos — ) A k + ( E sm — ) B k . 

0 &-»l ' l=»0 ^ > kwa> 1 ' »««0 


n / 


It can be established that (cf. Example 1, Sec. 17, Chap. 2) 


and 

Therefore, 


2 n — 1 

ikir 


E 

cos — 

= o , 

t ~0 

ft 


2 ft — 1 

ikir 


r 

sin — 

= 0 , 

i «*0 

ft 



2n — ] 

2n A o ~ 2 2/*’ 


i « 1,2, ...,» 

A: = 1, 2, — 1 


(12-3 


Multiplying both sides of each equation in (12-2) by the coefficient of /. 



713 


SBC. 12] INTERPOLATION AND EMPIRICAL FORMULAS 

in it, and adding the results, give® 

•JK 1 yV " Z 2 ^ 1 ik* ijr\ 

2^ Vi cos — « 2_ ( 2L> cos — cos — 14* 

i—0 W A:«»«l \ ta»0 ^ W / 

-- 1 Z 2 ^ 1 . *r $w\ 

+ 2* l 2^ sm — cos — 1 Bk 

. . , _ . Jb— i \ *~o n n/ 

for j * 1, 2, . * n — 1. But 


2 ^ 1 ikv ijir 

2j COS COS = 0, if fc 5*i j, 

*—0 


n 


n 




if k = j, 


and 


2 ” 1 . ifcir ijv 

2^ sm — cos — = 0 
t~ o n n 


for all values of k. Therefore, 

2n- 1 

nAy = 23 2/* cos 

.-o n 


^7T 


j = 1, 2, .. n - 1. 


(12-4) 


To determine the coefficient of A n the procedure is precisely the same, 
but 

2 lZ l ikir 

> , cos — cos zV »: 0, if k n, 

*-o n 

= 2n, if & = n. 


Hence, 


2w-1 

2nA n * 23 2/» cos **• 

»=»o 


(12-5) 


Similarly, on multiplying both sides of each equation of (12-2) by the 
coefficient of Bk in it and adding, one finds that 

2 1^ 1 ijir 

nBj = 2 j/, sin — . j = 1, 2, . . ., n - 1. (12-6) 

.-o n 

Equations (12-3) to (12-6) give the constants in (12-1). A compact 
schematic arrangement is often used to simplify the labor of evaluating 
these constants. It will be illustrated in the so-called “6-ordinate” case, 
that is, when 2n = 6. The method is based on the equations that deter- 
mine the constants, together with relations such as 

ir (n — 1)t (n + 1 ) 7T (2 n — 1)tt 

sin - ® sin * — sin = — sin » 

n n n n 


ir 

cos - » 
n 


(n — 1 )tt 

— cos 

n 


— cos - 


(n + I)** 


cos 


(2n — l)ir 


n 



714 NUMERICAL ANALYSIS [CHAP. 9 

Six-ordinate Scheme . Here, 2n * 6; the given points are where 

x, » mt/ 3 (i * 0, 1, 2, 3, 4, 5) ; and Eq. (12-1) becomes 

y « Aq + ili cos x + A 2 cos 2x + A 3 cos 3x + Bi sin x + B% sin 2x. 

Make the following table of definitions: 


yo y\ y 2 

Vo Vi 

V)o Wi 

Vz Vi 2/6 

H 

Wi 

Sum . . Vo Vi 

Vo Pi 

ro n 

Difference (iz?o w% 

Qi 

*1 


It can be checked easily that Eqs. (12-3) to (12-6), with n = 3, become 
&4o - Vo + Pu SAi * r 0 + ^s 1? 3^4 2 = Po — 

VS VS 

6^3 * r 0 - s h ZB\ « — r b 3£ 2 = — <?i- 

2 2 

Example: In particular, suppose that the given points are 


X 

1 

0 

7T 

3 

2 tt 

3 

V 

4r 

3 

Sir 

3~ 

2ir 

y 

1.0 

1.4 

1.9 

1 7 

1.5 

1.2 

1.0 


Upon using these values of y in the table of definitions above, 


1.0 


1.4 

1.9 

1.7 


1.5 

1.2 

2.7 

Vi - 

2.9 

t *2 ** 3. 1 

- 0.7 

* 

- 0.1 

u * * 0.7 


2.7 


2.9 

- 0.7 

- 0.1 



3.1 


0.7 

po « 2 7 

Pi M 
?i - 

6.0 
-0 2 

ro - — 0.7 

n * 0.6 

*1 * — 0.8 


Therefore, the equations determining the values of the constants are 


64 0 * 2.7 + 6.0 - 

8.7 

and 

a 0 - 

1.45, 

3A] ■» —0.7 —04 = 

-1 1 

and 

A 1 - 

-0.37, 

3 ^ 2 « 2.7 - 3.0 * 

-0.3 

and 

At — 

-0.10, 

6A 3 * -0.7 + 0.8 - 

0.1 

and 

A, - 

0.02, 

Vi 

3Bi - — (0.6) - 

jS 

0.3V3 

and 

B, - 

0.17, 

Vi 

3Bj — - 0.2) - 

m 

-0.1 V3 

and 

Bj “ 

-0.06. 



SEC. 13] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 715 
Hence, the curve of type (12-1) that fits the given data is 

y — 1.45 — 0.37 cos x — 0.10 cos 2x + 0.02 cos 3x 4* 0.17 sin x ~ 0.06 sin 2x. 

A convenient check upon the computations is furnished by the relations 

Vi 

Ao 4- + -^2 4* ^4.8 *■ t/o and #i 4- B 2 ** (yi — Fs)» 

Substituting the values found above in the left-hand members gives 

1.45 - 0.37 - 0.10 4 0.02 - 1.0 and 0.17 - 0.06 - 0.11, 

which check with the values of the right-hand members. 

Similar tables can be const ructed for 8-ordinates, 12-ordinates, etc. 


NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 


13. Numerical Integration. ^ The reader is familiar with the interpreta- 
tion of the definite integral / /(a) dx as the area under the curve y = /(a) 

j a 

between the ordinates x « a and x « b. This interpretation underlies the 
construction of formulas for numerical integration contained in this sec- 
tion. 

It will be recalled that if the function f(x) is such that its indefinite 
integral can be obtained, then the fundamental theorem of integral calculus 
provides an easy means for evaluating the definite integral. 1 However 
when /(a*) does not have an indefinite integral expressible in terms of known 
functions, or w r hen the values of /(a) are given in tabular form, formulas 
for numerical integration are generally used to obtain an approximate 
value of the integral. 

Formulas for numerical integration, or mechanical quadrature , are ob- 
tained by replacing the function /(a) specified at a given number of points 
in the interval (a,b) by a polynomial (8-5) or (9-2), depending on whether 
the values of a are equally or unequally spaced. 

If the values of y - /(a) are known at m 4 1 points x x , where i = 0, 
1, 2, . . . , m, which are spaced h units apart, an approximate value of the 

integral / m /(a) dx can be computed by substituting in the integrand an 
Jzo 

approximate polynomial representation of y = /(a) given by (8-5) or, 
equivalently, (8-6). We thus get for equally spaced values a* 

£y dX = JT [vo + X Ay 0 + 


H h 


X(X - 1) ... (X - m + 1) 

to ! 



dX, 


(13-1) 


1 Bee Chap. 3, Sec. 13. The evaluation of difficult integrals by power series is discussed 
to Chap, 2, Sec. 10. 



{chap. 9 


716 NUMERICAL ANALYSIS 

where X is the dimensionless variable defined by 


x — x 0 


and X = to for 


X m = x 0 + mh. 

If to * 1, formula (13-1) yields 


(13-2) 

(13-3) 


Aj/o Vi - Vo 1 


J Q VdX **f o (y 0 +X A Vo) dX * y 0 + = J/o + 2 


" (ifo + yi). 


But from (13-2) dX « rfar/A, and on recalling (13-3), we see that this 
formula can be written as 

/ l ydx = ~ ( 2/0 + Vi). (13-4) 

-'xo 2 

Since yo is the ordinate of y = /(x) at x = x 0 and 2 / x is the ordinate at 
x » Xj, the right-hand member in (13-4) represents the area of the first 
tr&peaoid sliown in Fig. 13. The choice of m = 1 in the calculations 



leading to (13-4) corresponds to replacing y = f(x) in the interval (x 0 ,x 1 ) 
by the straight line through ( 20 , 2 / 0 ) anc ^ ( x ifVi)- 
The successive application of (13-4) to intervals (xi,x 2 ), ( 22 , 23 ), . 
(x n -i,x n ) yields 


L* y dx 


B [ \dx + f *ydz+' 

Jxq Jx 1 


■+r v dx 

Jxn-\ 


** - (2/0 + 2/1) + - (yi + 2/2) H t- - (2/n — 1 4 - y«) 

& Z 2 

a , 

— - (l/o + 2yi + 2 j/ 2 + • • • + 2y n -\ + y n )- 


(13-6) 



SEC. IS] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 717 

Formula (13-6) is known as the trapezoidal rule, for it gives the value of 
the sum of the areas of the n trapezoids whose bases are the ordinates 
Vot VuVat • • * » Vn- Figure 13 shows the six trapezoids in the case of n = 6. 
If m - 2, (13-1) becomes 

r a f 2 f (X 2 - X) 0 1 

j^ydX = jf 1 2 /o + A 2 y 0 \ dX 

1 /8 \„ 

= 2y 0 + 2 Ay 0 + - (- - 2 ) A 2 y 0 
1 

= 2y 0 + 2(yi - y 0 ) + ~ (ya - 2y l + Vo) 

O 

1 4 1 

= - Vo + ~Vi + ~Va, 

r*t h 

or / y dx * - (y 0 + 4y x + y 2 ). (13-6) 

■ / aro 3 

Suppose that there are n + 1 pairs of given values, where n is even. If 
these n + 1 pairs are divided into the groups of three pairs with abscissas 
# 2 t> # 2 »+ 2 > where i = 0, 1, . . . , (n — 2)/2, then (13-6) can be applied 

to each group. Hence, 

[ x n [ x i f x 4 f x n 

/ y dx ~ I ydx+ y dx d b / y dx 

•''*0 ^xa Jx n ~i 


h h 

* - («/o + 42/1 + 2/2) + - (l /2 + 4j/3 + 2 / 4 ) 

u u 


H b ~ iVn —2 + 4y n -i + 2/n) 

u 

h 

“ - [j/O + 2/n + 4(2/1 +2/3-1 1- 2/n-l) 

o 


+ 2(y 2 + 2/4 H — * + Vn- 2 )]- (13-7) 

Formula (13-7) is known as Simpson's rule with m =® 2. Interpreted 
geometrically, it gives the value of the sum of the areas under the second- 
degree parabolas that have been passed through the points ($ 2 ^ 2 %) , 
(*w+idto+i)> (*ai+ 2 ,ya<+a)» where i « 0, 1, 2, • • « > (n — 2)/2. 



718 


NUMERICAL ANALYSIS 


{char. 9 


If m ** 3, (13-1) states that 

X s - X , X s - 3X 2 + 2X , \ „ 

J 0 VdX - J o yy 0 + X A y 0 H — A 2 j/ 0 -I Al/oJ dX 

9 (9 9\ /27 9 3\ , 

= 3j/o + - Aj/o +(--J A J/o + (- - - + -j A j/o 

9 9 

“ 3j/o + - (Vx - 2/o) + 7 ( 2/2 - 2yi + j/ 0 ) 

2 4 

3 

+ ~ (Vs — 3^/2 + 32/i — 2/o) 

o 

3 

=*“(2/0 + 3t/i + 32/2 + 2/3), 

o 

r* , 3/i 

or / y dx « — (2/0 + 3z/i + 32/2 + 2/3). (13-8) 

•^o 8 


If n + 1 pairs of values are given, and if n is a multiple of 3, then (13-8) 
can be applied successively to groups of four pairs of values to give 

£ n 3 h 

y dx ~ [ 2/0 + y n + 3(2/1 + V2 + V\ + 2/5 4 + Vn—2 + Vn- 1 ) 

8 

+ 2(2/3 + 2/e H 1- Vn- a)]- (13-9) 

Formula (13-9) is called Simpson's rule with m = 3. It is not en- 

countered so frequently as (13-5) or (13-7). Other formulas for numerical 
integration can be derived by setting m = 4, 5, ... in (13-1), but the three 

given here are sufficient for ordinary purposes. In most cases, better 

results are obtained by securing a large number of observed or computed 
values, so that h will be small, and using (13-5) or (13-7). 


Example 1. Using the data given in the Example of Sec. 7, find an approximate value for 

J^ydx. 

Using the trapezoidal rule (13-5) gives 

J^ydx - HC2.105 4~ 5.616 4- 7.228 -f 9.208 -f 11.714 4- 14.902 + 9.467) » 30.120. 


Using (13-7) gives 

J\dx - H12.105 -b 9.467 + 4(2.808 4- 4.604 4- 7.451) 4- 2(3.614 4* 5.857)] « 29.989. 
Using (13-9) gives 

f ydx - H\2.m 4 * 9.467 4- 3(2.808 -f 3.614 4- 5.857 4- 7.451) 4- 2(4.604)] - 29.989. 



13] NUMERICAL INTEGRATION OF OlFFERENTIAL EQUATIONS 719 

If numerical integration is to be used in a problem in which the form of 
f(x) is known, the set of values (xi,y t ) can usually be chosen so that the 
Xi form an arithmetic progression and one of the formulas deduced above 
can be applied. Even if it is expedient to choose values closer together 
for some parts of the range than for other parts, these formulas can be 
applied successively, with appropriate values of h , to those sets of values 
for which the x t form an arithmetic progression. However, if the set of 
given values was obtained by observation, it is frequently convenient to 
use a formula that does not require that the x x form an arithmetic progres- 
sion. 

Suppose that a set of pairs of observed values ( Xi t y x ), where i = 0, 1, 
2, . . . , m, is given. The points (x iy y x ) all lie on the curve whose equation 
is given by (9-2). The area under this curve between x ~ xo and x * x m 
is an approximation to the value of j y dx. The area under the curve 

J x o 

(9-2) is 

/*" y dx — £ ~~ / " Pk(x) dx, (13-10) 

Jx ° Pk(Xk) Jxt 

in which the expressions for the P*(x) are given by (9-1). 

If m « 1, (9-1) and (13-10) give 

f * y dx - — — — [*' ( x — x x ) dx 4 — — f* 1 (x — rr 0 ) dx 

J x Q X q Xi •'*0 X\ ~ Xq 

X\ — Xq 

= — (2/o + 2/i). (13-11) 

A 

Formula (13-11) is identical with (13-4), as would be expected, but the 
formula corresponding to (13-5) is 

/ ydx = l A[(xi - X 0 )(y 0 + Vi) + (*2 - Xi)(yi -f y 2 ) 

J x 0 

H h (x„ - x n _,)(y n _! + y n )]. (13-12) 


If m = 2, (13-10) becomes 



720 


NUMERICAL ANALYSIS 


[chap. 9 


Vo r j - Xq _ (*l + X 2 )(x 2 - Jo) 

Po(*q) L 3 2 


+ XiX2(%2 


-*#)] 


vi r 
M*i) L 


x\ - xl (x 0 + X 2 ) (xl - To) 


Va \x\~xl (x 0 + x,)(xf - Xq) 


P*{*a) L 

(*2 ~ *o) 2 f Vo 


+ XoX 2 (xa 


+ XoXi(x 2 


-*.)] 

~ *<>)] 


— - (3x 1 - 2x 0 - x 2 ) + (x 0 - x 2 ) 

(x 0 ) Pl(Xl) 


+ (2x 2 + Xo - 3xj) • (13-13) 

PaM J 

Formula (13-13) reduces to (13-6) when X\ — £ 0 ~ x 2 — 'Xi = h . The 
formula that corresponds to (13-7) is too Jong and complicated to be of 
practical importance, and hence it is omitted here. It is simpler to apply 
(13-13) successively to groups of three values and then add the results. 

Example 2. Using the data given in Prob. 3, Sec. 9, find an approximate value of 


r .2ft 

ydx . 

it 


Using (13-12) determines 


r 0.26 

/ ydx - ^(0.24(4.210) + 0.6(4.631) + 1.5(5.082) + 3.75(5.590)] - 16.187. 

JO At 

Applying (13-13) successively to the first three values and to the last three values gives 

r M dx m (0,84 )* r 2 ( L2 ~ °' 32 ~ 2 - 21 °(-°- 84 ) 2.421(2 +0.16 - 1.2) *1 

H V “el (~0.24)(— 0.84) + (0.24)( — 0 6) + (0.84)(0.6) ~J 


(0.84)(0.6) 


(5.25) 2 ["2.421(7.5 - 2 - 6.25) 2 601 (-5.25) 


(-1.5)(-5.25) 


(1.5)(— 3.75) 

2.929(12.5 + 1 - 7.5)" 
+ (5.25) (3.75) 


PROBLEMS 


Determine the values of ydx by applying (13-5) and (13-7) to the following data : 


xl 234567 
y 2.157 3.519 4.198 4.539 4.708 4.792 4.835 



SBC. 14 ] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 721 


1 . 50.645 

2. Apply formula (13-12) to compute / p dv from the data of the Example in Sec. 9. 

h o 

3. Work the preceding problem by applying (13-13). 

>25 

4. Apply formulas (13-5) and (13-7) to compute / H dC from the data of Prob. 4, 

Sec. 9. . 19 

5. Find approximate values of j V4 4- x 3 dx by applying formulas (13-5) and (13-7) 

with Xm *® tn, m » 0, 1, 2, . . 6. 


14. Euler's Polygonal Curves. The methods available for the exact 
solution of differential equations, as we noted in Chap. 1, apply only to a 
few, principally linear, types of differential equations. Many equations 
arising in applications are not solvable by such methods, and one is obliged 
to devise techniques for the determination of approximate solutions. 

We begin with the consideration of the first-order equation 

V' = f(x,y) ( 14 - 1 ) 


and seek its solution y = y{x) taking on a prescribed value yo = y(x 0 ) at 
x « z 0 . 

At each point of the region where f(x,y) is continuous Eq. (14-1) deter- 
mines the slope of the integral curve passing through that point. The 
equation of the tangent line at the point (x 0> y Q ) to the integral curve 
V = v(x) is 

y - y 0 = f(x 0} y 0 )(x - So). (14-2) 


If we advance along this line a short distance to a point (21,2/1), we can 
compute from (14-1) the value y\x x ) = f(x X) y x ) which, in general, will 
not be equal to the slope of y = y(x) at x = x lt because the point (xi,^) 
ordinarily will not lie on the integral curve y = y(x). But if (x x ,y x ) is 
close to (xo,yo), the slope of the integral curve at x = 2! will not differ 
much from f(x X) y x ). To put it differently, the linear function (14-2) 
approximates the solution of (14-1) in the neighborhood of the point 
(ar 0 ,t/o)* (^ ee Fig. 1 i n Chap. 1, Sec. 1.) 

We consider next the straight line through (x x ,y x ) with the slope f(x Xl y x ) 
and proceed along it a small distance to a point (22,2/2). At (22,2/2) we 
draw another straight line with the slope /( 22,^2) an( * advance along it to 
a point (23,2/3). By continuing this construction we obtain a polygon 
consisting of short straight-line segments joining the points (20,^0), (21,2/1), 
(22,2/2), •••, (2 n ,$/n)* The polygonal curve so obtained is called Euler’s 
polygon . This polygonal curve can be expected to approximate the in- 
tegral curve reasonably well when the points (x*,2/ t ) are not too far apart 
and the end point (x n ,Vn) is not too far away from (20,2/0). 

The end points of the segments forming Euler’s polygon clearly satisfy 



722 NUMERICAL ANALYSIS [CHAP, 9 

[of, (14*2)) 

Vi ~ Vo * /(®o,Jfo)(*i - *o) 

1/2 - 2/1 = f(x\,V\)(x2 ~ X\) (14-3) 


|/n - 2/n— 1 = /(^n-l,2/n~l)(^n - «n-l) 

and if each interval x % — x x _>i is of length h } we can write (14-3) as 

2/m+l 5=5 2/m + f(Xm,ym)h, ™ = 0, 1 , 2, . . . , U - L (14-4) 

The recursion formula (14-4) enables us to compute successively the 
approximate values of the ordinates of the integral curve y = y{x) at 
Xk * Xo + kh, where A; * 1, 2, . . . , n. It may suffice for rough calculations 
if the spacing interval h is small and m not too large, 

A more accurate formula can be obtained by constructing, instead ^f 
the chain of rectilinear segments, a chain made up of parabolic segments. 
Thus, we can draw through (xi,yi) a parabola 

y «■ ao + a x (x — x{) 4- a 2 (x — x x ) 2 (14-5) 

which at x = Xo has the slope /(xo,2/o) and at x = x 3 the slope f{x\ } yi). 
A simple calculation of the constants in (14-5) yields 

2/2*2 /i + {y'(xx) + l A[y'(x x ) - y'(x 0 )]}h. (14-6) 

This formula serves to determine y 2 if 2/i =» y(x i), y'(xi), and the difference 
Vy[ ss y'(x x ) — y'(xo) are known. Now, if we suppose that the solution 
y(x) can be represented by Taylor’s formula 

y(z) <= i/o + y'(xo){ x - x 0 ) + Hy"(xo)(x - x 0 ) 2 H h R„, (14-7) 

we can calculate the needed quantities in the right-hand member of (14-6). 

The coefficients in (14-7) can be calculated from (14-1) whenever f(x,y) 
has a sufficient number of partial derivatives, for on setting (x 0 ,2/o) in 
(14-1), we get j/'(xo) «/(xo,2/o). Differentiating (14-1) with respect to 
x yields 1 

y”{x) ~ f x (x t y) + /„(*, y)y'(x) 9 (14-8) 

and substituting x * Xo, y - 2/o in (14-8) gives 

J/"(*o) * fx(x 0l y 0 ) + f v (xo,y 0 )y'(x Q ). 

By differentiating (14-8), we obtain t/"'(x), and so on. The value of R n 
in (14-7) in general caimot be computed, but by neglecting it we get an 
approximate value of y(x ). 

Once the coefficients in (14-7) are determined, we use (14-7) to compute 
1 We use the subscript notation for partial derivatives introduced in Sec. 2, Chap. 3* 



SEC. 15] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 723 

y(x i) = yi. The value of y'(x i) is then determined by (14-1), since y'(z x ) 
* f(x lt yx). The substitution in (14-6) then yields y 2 . 

Having computed y 2} we can advance another step and compute t/ 3 
from [cf. (14-6)] 

yz = V* + \y'{x 2 ) + YzWixi) - v'(^i))} A- 

This requires calculating y f (x 2 ) from (14-1). 

The general recursion formula, based on the parabolic approximation, is 

2/m 4-1 “ 2/m + Iv'&m) + Vl Vy'(x m )]/l, (14-9) 

where Vy'(x m ) « y'(x m ) - y'(x m _i). 

More elaborate recursion formulas can be constructed by using poly- 
nomials of higher degree instead of (14-5). Such formulas lie at the basis 
of the Adams method of integration of differential equations discussed in 
the next section. 


PROBLEMS 

1. Construct a polygonal approximation, in the interval ( — 1,1), to the solution of 
y* «■ %xy which is such that y( 0) * 1. Take the spacing interval h * 0.2. Also obtain 
the exact solution, and plot it on the same sheet of paper. 

2 . Determine the coefficients in (14-5), and thus deduce (14-6). 

S. Use the equation in Prob. 1 to illustrate the calculation of y<i from formula (14-6). 
Also obtain y % . Take xo *= 0, yo * 1, a*i * 0.2, x% * 0.4. 

16. The Adams Method. We extend the considerations of the preceding 
section by developing a step-by-step procedure for computing an approxi- 
mate solution of 

y f = f(z,y) (15-1) 

taking on a prescribed value yo at x = To. The ordinates y my approximat- 
ing the ordinates of the integral curve y = y{x) at z — z TO , will be de- 
termined for equally spaced values of z, so that x m = Zo + hm f where 
m = 0, 1, 2, Thus our approximate solution will appear in a tabu- 

lated form for a discrete set of values of z. 

By the Fundamental Theorem of Integral Calculus, 

r~ + ' y’(x) dx = f n+ ' (y\ dx = y(x m + h) - y(x m ) 

J *m Jx m \dx/ 

bo that Vf+i = Vm + [ " + * y'(x) dx, (16-2) 

* xm 

where m y(x m + h) and y m « y(x m ). 

Now, if the variable x in the integral of (15-2) is replaced by 

* = x m + hX (15-3) 



724 NUMERICAL ANALYSIS [CHAP 9 

where Jf is a new dimensionless variable, (16-2) becomes 1 

y m +i - y m + h j[‘ y'(x m + hX) dX. (16-4) 


But we saw in Sec. 8 that when a function y'(x) is approximated by a 
polynomial of degree n taking on the values y m , y„-i, . . .,y m -„ at x = x m , 
x n -t, x m _„, then [cf. (8-7)] 


, , X(X + 1 ) , , 

tfizn + hX) = y' m + X Vy’ n + — — < -v*y m + --- 

« ! 


X(X + 1) . . . (X + n - 1) 

+ — — r -vVm- (15-5) 


If we insert (15-5) in (15-4) and carry out simple integrations, we find that 

Vm +1 ~y m + h(y m + H + V\i Vy m + H ^y m + 25 K 2 o 

+ ■■■+ a n V n y' m ), (15-0) 


where 


a n = / 

Jo 


i X(X + 1) . . . (X + n - 1) 


n\ 


dX, 


(15-7) 


Formula (15-6) enables us to compute the ordinate y m +\ if we know 
y m , y and the backward differences V k y' m . When the Vy m vanish, (15-6) 
reduces to (14-4), and when the V 2 y' m vanish, we get (14-9). As was the 
case with (14-9), the values of y m , y m and the V k y m in (15-6) are not avail- 
able to us at the start. They must be computed by some means before 
(15-6) can be used to evaluate y M +\. The number of the V k y m depends on 
the degree n of the polynomial chosen to approximate y'(x). Once we 
agree on the value of n, we can compute y m , y m and the requisite number 
of the V k y'm with the aid of Taylor’s representation of the solution y — y(x), 
as was done in Sec. 14. 

We illustrate the procedure in detail in the following example. 

Example : Use Adams’ method to obtain, in the interval (0,1), an ap- 
proximate solution of 

y’ * y + x, (15-8) 

taking on the value y$ * 1 at x = 0. 

Let us subdivide the interval (0,1) into subintervals of length h * 0.1, 
so that 

Xk *= %o + kh ~ 0.1/c, k = 0, 1, 2, . . 10. 

Furthermore, let us agree to retain in (15-6) the differences of y m up to and 
including those of order 3. This corresponds to approximating y f {x) in 
(15-2) by a polynomial of degree 3. 


1 By (1M) dx**h dX , and at the limits x » Xm and x • Xm -f h, the values of X 
are X * 0 and X « 1. 



SBC. 15] NUMERICAL INTEGRATION OF UIFFERENTIAL EQUATIONS 725 

To compute y m +i from (15-6) we need y m y m , Vy m V 2 y m) and V z y m . The 
calculation of the third differences V^y f m requires at least four values 
y'mi Vm~i) Vm- 2? Vm^h as obvious from the following table. 


Vm~Z 

Vm-2 

VVm-2 

V 2 Vm-l 


Vi/m-1 

V»y- 

y'm-i 


V 2 y' m 

Vm 



If we determine y' 0 , y\, y 2 , 2/3, we shall be in a position to fill in the values 
in this table with m — 3 and then proceed to determine y\ from (15-6). 
Since ?/o = 1 for rr 0 - 0, Eq. (15-8) yields 

Vo = 1- (15*9) 

To compute y j, y 2 , and 2/3 we use Taylor’s series 

n ftt 

y(r) = 2/o + 2/0(2; - x 0 ) + ^ (x - x 0 ) 2 + ~ (x - x 0 ) 3 + • • • (15-10) 


with 2/0 — 1 and .r 0 — 0. The coefficients in (15-10) can be calculated 
irom (15-8). Differentiating (15-8), we get 

y'(j) = y'(x) + 1, (15-11) 

and on setting x — 0 and recalling that 2/'(0) =2/0=1, we get y"(0) * 2. 
Successive differentiations of (15-11) give 

y"\x) = y" (x), ?/ v (x) = 2/"'(*), . . J/ (n) (x) = y^fc), 

(15-12) 


and since 2/"(0) = 2, we get from (15-12) 

y"'(0) = 2, 2/ IV (0) - 2, 

Accordingly, (15-10) becomes 


x 3 x* T S 

2-1 I | 


y(x) = 1 + i + r H b 

3 3-4 3-4-5 

Setting x = 0.1, we get 

, ( 0 . 1) 3 ( 0 . 1) 4 ( 0 . 1) 6 

2/i = 1 + 0.1 + (0.1) 2 + + i— + —— + 

o o ’ 4 o 1 4 * 0 

In the same way using x = 0.2 and x = 0.3, we obtain 
l/a - 1.2428, y 3 - 1.3997. 


2/ (n) (0) = 2. 
+ •••• 


1.1103. 



NUMERICAL ANALYSIS 


726 

The desired values of y\, y 2) and y$ 
We find that 


[chap, 9 

can now be computed from (15*8). 


Vi - Vi + xi « 1.1103 + 0,1 « 1.2103 

^2 ** V2 + x% =» 1.2428 + 0.2 = 1.4428 

Vz 385 Vz + x z “ 1.3997 + 0.3 » 1.6997. 

We can now proceed to construct the table of differences shown below. 


X 

y 

y f 

W 

vV 

vy 

0 

1.0000 

1.0000 







0.2103 



0.1 

1.1103 

1.2103 


0.0222 





0.2325 


0.0022 

0.2 

1,2428 

1.4428 


0.0244 





0.2569 


| 0 0026 

0.3 

1 3997 

1.6997 


0.0270 

i 




0.2839 



0.4 

1.5836 

1.9830 




0.5 

1.7974 






The substitution from this table in (15-6), with m = 3 and n = 3, yields 
y A - 1.3997 + 0.1(1.6997 + H(0.2569) + Ha(0.0244) + %(0.0022)] 

« 1.5836. 

This value is recorded in the table for x = 0.4. 

To compute we must extend the table, since formula (15-6) requires 
the knowledge of y\ and assorted differences of y\. By (15-8) 

y\ * y 4 + * 1.5836 + 0.4 = 1.9836. 

The calculated values (recorded below the heavy line in the table) can 
now be used in (15-6), with m = 4, n = 3, to compute y 6 . We have 

y 5 - 1.5836 + 0.1(1.9836 + M(0.2839) + ^2(0.0270) + %(0.0026)J 

* 1.7974. 

This value is recorded in the table for x — 0.5. 

We leave it to the reader to make further extensions in the table re- 
quired for the calculation of y yr, . . . , yio- 


PROBLEMS 

1. Complete the table in the Example of Sec. 15 by computing yt> y 7, ...» y\Q. 

2. Since (15-8) is a linear equation, its solution satisfying the condition y(0) — 1 



EEC. 16] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 727 

is easily found to be y * 2e* — x — 1. Compare the exact and the approximate values 
Vh V* " * Vw- 

3. Apply the Adams method to obtain an approximate solution of y* « y with y(Q) 
« 1. Use h — 0.1, and compute y( 0.3), jy(0.4), y( 0.5), and ?/(0.6) from (16-6) with n «* 2. 
Compare with the exact solution. 

4. Use k « 0.1 and (15-6) with n « 3 to find an approximate value of y(~0.6) for 
the integral curve of y’ « x 2 4* y 1 through (—1,0). 


16, Equations of Higher Order. Systems of Equations. The methods 
of Secs. 14 and 15 can be extended to obtain numerical solutions of equa- 
tions of higher order. Thus, the second-order equation 

v" = f(z,v>y') (16-1) 

with initial conditions 

y(* o) = Vo, y'{x o) = y 0 (16-2) 

can be written as a system of two equations of first order by setting 

y' = z. (16-3) 

The substitution in (16-1) from (16-3) then yields the second equation 

2 ' =/(*,!/, 2 ). (16-4) 


In indicating the extension we shall consider, instead of the system (16-3) 
and (16-4), a more general system 

y' = fi(x,y,z), 


2 ' = f 2 ^,V,z), 


(16-5) 


with initial conditions 

y{x Q ) = 1 / 0 , 2 (x 0 ) = z 0 . (16-6) 

When solutions of the system (16-5) can be expanded in Taylor’s series 


y(x) = y(x 0 ) + y'(x 0 )(x - x 0 ) + ----- (x - x 0 ) 2 + • 

Z ; 

z(x) = z(j q) + z'(x 0 ) ( x - * 0 ) + (x - x 0 ) 2 + • 


(16-7) 


the coefficients in (16-7) can be computed by differentiating Eqs. (16-5) 
successively as was done 1 in Secs. 14 arid 15. 

The construction of Euler’s polygonal approximation also follows the 
pattern of Sec. 14. Thus, the equation of the straight line through (xo, Vo,Zq) 
tangent to the integral curve of the system (16-5) is 2 


V - V o * /i(*o,2/o,Zo)(x - x 0 ), 
2 - So * - Xo). 


(16-8) 


1 Sec in this connection Sec. 6, Chap. 3. 

1 The integral curve of the system (16-5) is, in general, a space curve, so that the 
tangent line to it is determined by the intersection of the planes (16-8). 



728 NUMERICAL ANALYSIS (CHAP. 9 

When abscissas are spaced uniformly h units apart, 

#i ** #o + K x 2 ** x o + 2^, * . . , Xk ** xq + kh, 
and from (16-8) it follows that the approximate solutions at x Xj x 3 , * . . are 
Pi “ Po + fi&o, PQ,Zo)h, 

Zi « Zo +/2(zo,2/o,zo)fc, 

P2 - 2/1 

*2 = *1 + /»(*!, yi,«l)A, 


y*+i = 2 /a 

Z*+l * ** +/2(^Jfc,y/b,Zfc)^. 


If, instead of approximating the solution in each interval by a linear func- 
tion, we make use of the polynomial approximations in the manner of 
Sec. 15, we obtain 

Vm+t = y m + %■ + Vi Vy m + 5 /i2 Yy' m H h a„ TTy'j 

, , o , , (16-9) 

*m+l = 2m + h[z m + Y ^ z m H 1* a n ^Vm] 


with a n determined by (15-7). 

In computing y m +\ and z m +\ from (16-9), we must first obtain the values 
of Pm, Zm, y m , z m and the required differences, as was done in Sec. 15. 


Example: Obtain the solution of the system 

y' - x + * 


*' - l + y 


(16-10) 


in the form (16-7), which is such that 

1/(0) - -1, 2(0) ~ 1. 

On setting xo - 0 in (16-7) we get 

»(i) - 2,(0) + 2,'(0)i + ^ j/"(0)r 2 + • • • 

*(i) - z(0) + z'(0)x + ~ t"( 0)x 2 -) , 


(16-11) 


(16-12) 


the coefficients in which can be computed by differentiating (16-10) and noting (16-11). 
We obtain from (16-10) 


y"(x) « 1 + z'(x), 

«"(*> - v'M 

y"\x) - «"(*), 

t'"(x) - y"(x) 

y(*\x) ■■ * ( *“* 1} (x), 

z (n Hx) - V <— l >(x). 


(16-13) 



1 

SEC. 16] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 729 

The substitution from (16-11) in (16-10) yields y'(0) - 1, z'(0) - 1 — 1 - 0, and making 
use of these values in (16-13) we find 

V"( 0 ) -l+o-l, *"(0) - y'( 0) - l, 

V'"(0) - *"(0) - 1, *"'(0) - V "( 0) - 1, 

» <n, ( 0) - 1, #<»>( 0) - 1. 


Accordingly, (16-12) yields 

% ® xfi 

V(x) =-l + J + -+ -+... 

X ® 

z(x) — -f— n . 


(16-14) 


By eliminating z from the system (16-10) we see that it is equivalent to the second- 
order equation 

V" - y~ 2 

with y(0) — 1 f y'( 0) ™ 1. Its solution is readily found to be 

V * e* ~ 2, (16-15) 

and from the first of Eqs. (16-10) we conclude that 

z ® - x. (16-16) 

The Maclaurin expansions of these solutions are precisely (16-14). 

It may be instructive to compute the polygonal approximations to the solution of 
(16-10) ati * 0.2 and x — 0.4 
On setting the differences in (16-0) equal to zero, we get 

Vm+l ** Vm 4~ Ztn-fl c= *4" 

Now, if we take Xi ** 0.2, so that h * 0.2, we obtain from (16-17) 

Vi - y(0.2) - -1 + (0.2)1 - —0.8, 

zi * £(0.2) * 1 + (0.2)0 * 1, 

since yo 1 and zq «* 0. 

The exact solution (16-15) and (16-16) yields 

y(0.2) * c 0 - 2 - 2 - —0.7786, 

2(0.2) * e 0 - 2 - 0.2 * 1.0214. 

Using yi *» —0.8, «i * 1 in (16-17), we obtain 

Vi * y(0.4) « yi -f 0.2y{, 

z*t « 2(0.4) « 2i 4* 0.22i. 

The values of yi and z[ can be calculated from (16-10) by setting x « 0.2, 2 * 1 , and 
y » —0.8. We find that 

yi m 0.2 + 1 - 1.2, ti » 1 + - 1 - 0.8 - 0.2, 


(16-17) 


(16-18) 



730 

and then (1648) yield® 


NUMERICAL ANALYSIS 


[chap. 0 


y% « —0.8 4* (0.2) (1.2) » -0.56, 

2 2 - 1 4 (0.2)(0.2) - 1.04, 
while the corresponding exact values are 

y(0A) - e 0 * 4 - 2 - -0 5082, 

2(0.4) « c 0 * 4 - 0.4 * 1.0918. 

The reader is advised to obtain more accurate polygonal approximations by taking 
the interval h ** 0.1 and to compare the polygonal approximations with the values given 
by (16-9) in which the differences of order higher than 1 are set equal to zero. 


PROBLEMS 

1. Obtain from (16-7) a fourth-degree polynomial approximation to the solution of 

y f « e x -f 2, d ** 4 y 

with y( 0) *= 0 and z(0) «* 0. 

2 . Use a polygonal approximation to compute yi, y% yt, y* for the system in Prob. 1 

by taking x\ *» 0.1, *■ 0.2, x% — 0.3, x\ =* 0.4 

3. Use a polygonal approximation to compute yi, yz , corresponding to x\ *=0 1, 
X 2 — 0.2, xz ** 0.3 for y" — y 2 * x, with initial conditions y( 0) = 1, y'(0) » 0. fftnt 
Set « z, and consider the system y' = z, z' «* x 4 y 2 with j/(0) * 1 and 2(0) « 0. 

4 . Obtain the solution for Prob 3 in Maclaurin’s series. 

5. Solve the system in Prob. 3 by the Adams method. Retain only the second dif- 
ferences in (16-9), and use the result of Prob. 4 to start the iteration. 

17. Boundary-value Problems. In many physical problems solutions 
of the second- and higher-order differential equations are required which 
satisfy preassigned conditions at more than one point of the interval. A 
simple example of this occurs in the study of deflections of a beam supported 
at several points. Problems of this sort are termed boundary-value prob- 
lems to distinguish them from initial-value problems in which the conditions 
on solutions are imposed only at one point. 

An important feature of the boundary-value problems is that their 
solutions (if they exist at all) need not be unique . 1 When the general 
solution of the differential equation can be obtained, the conditions im- 
posed on solutions of the boundary-value problem can usually be met by 
determining the values of arbitrary constants in the general solution 2 
so that the specified conditions are satisfied. How r ever, general solutions 
of differential equations can rarely be written down, and one is obliged to 
seek solutions of boundary-value problems by numerical methods. The 

1 See, for example, our discussion of two interesting two-point boundary-value prob- 
lems in Sec. 34, Chap. 1. 

*This was the procedure followed in solving the boundary-value problems in Sec. 
34, Chap. 1. 



1 

SEC. 18] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 731 

methods available for numerical solution of initial-value problems require 
that the integral curve be uniquely determined at the starting point and 
thus do not apply to problems in which solutions must satisfy specified 
conditions at more than one point. To solve a boundary- value problem 
numerically one must employ laborious trial-and-error procedures utilizing 
the solutions of suitable initial-value problems. 

We outline briefly the procedure commonly followed in solving a two- 
point boundary-value problem for the second-order differential equation. 1 

Let it be required to determine a solution of 

y" * f(z,y,y') (17-1) 

which assumes at the end points of the interval a < x < b the values 

y(a) = A, y(b) « B. (17-2) 

Now, if in addition to the value y(a) = A we specify the slope y'(a) 
at x = a, the solution of (17-1) is uniquely determined, 2 but this solution 
will satisfy the condition y(b) = B only for some value of the slope y'(a) 
which is not known. 3 Physical or geometric considerations may suggest 
an approximate value of the slope, say y'(a ) = C, which is such that the 
integral curve of (17-1) satisfying the conditions 

yip) = A, y'(a) = C (17-3) 

also satisfies the condition y(b) ~ B. 

The procedure used in solving the boundary-value problem consists in 
actually constructing the solution y = y(x) satisfying the conditions (17-3) 
and computing the value of y{x) at x = b. If it is tolerably near B, w r e 
have the desired approximate solution of the boundary- value problem. 
If not, we choose another value of the slope y'(a) and try again. The 
procedure is clearly laborious and far from being elegant. 

18. Characteristic-value Problems. Closely associated with boundary- 
value problems are charartcristic-value problems. These are generally 
concerned with solutions of the two-point boundary-value problems for 
differential equations containing parameters. 

A simple instance of the characteristic- value problem occurs in the study 
of small vibrations of an elastic string of finite length. 4 When initial shape 
and initial velocity of the string are specified, its subsequent displacement 

1 For a more detailed discussion of such problems see W. E. Milne, “Numerical 
Solutions of Differential Equations,” chap. 7, John Wiley & Sons, Inc., New York, 
1952. 

8 We suppose that f(x,y,y') is such that the initial-value problem has a unique solution. 

8 We assume that the boundary- value problem in (17-1) and (17-2) indeed has a 
solution. 

4 See Chap. 6. 



NUMERICAL ANALYSIS 


[chap. 9 


(18-1) 


are fixed at x 
conditions 


u(x f t) is determined by solving the equation 

d 2 u d 2 u 

where a is a physical constant. If the string is of length l and its ends 
0 and x * l, the solution of (18-1) must satisfy the end 

w(0,0 «. 0, u(Jt,t) « 0. (18-2) 

When we attempt to obtain solutions of (18-1) by the method of separa- 
tion of variables, 1 that is, by assuming that u{x,t) is expressible in the form 

» y{x)T(t), (18-3) 

where y(x) is a function of x alone and Tit) is a function of t alone, we are 
led to a pair of ordinary differential equations 

d 2 y 

+ X 2 y - 0, (18-4) 


dx 2 
d 2 T 


0 , 


where X is a constant. This constant must be chosen so that the end 
conditions (18-2) are satisfied. 

From the assumed form of solution (18-3) and from (18-2) it follows 
that the solutions of (18-4) must be such that 


y( 0) - o, y(l) « 0. (18-5) 

We thus have a two-point boundary-value problem for Eq. (18-4) with the 
end conditions (18-5). 

The determination of suitable solutions this time is very simple because 
the general solution of (18-4) is 

y « ci cos Xx + c 2 sin \x. (18-6) 

If we impose the conditions (18-5) on (18-6) and reject the trivial solution 
y « 0, we find infinitely many solutions 

y * c 2 sin \x, (IB-7) 

kv 

where X = — . k *1,2,..., (18-8) 

The values of X in (18-8) are called the characteristic values of the boundary- 
value problem of (18-4) and (18-5), and the solutions (18-7) with appro- 
priate Xs are characteristic functions of this problem.* 

1 See Sec. 10, Chap. 6. 

* The terms eigenvalue and eigenfunction are used by some writers to mean “charac- 
teristic value” and “characteristic function,” respectively. These stem from German 
words Eigemwert and Eigenfunktion, We eschew the hybrids, since this book is written 
in English. 



SEC. 18 ] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 733 

The simplicity erf the characteristic-value problem defined by (18-4) 
and (18-5) masks some important features of the general problem- These 
features become clearer if we consider the determination of small vibrations 
of an elastic rod of variable cross section. In this case the separation of 
variables in the appropriate partial differential equation leads to the equa- 
tion 

d 2 r d 2 yl 

fo? L ^ dz * J ~ * ° f ( 18 ** 9 ) 

in which p(x) and q(x) are known functions and X an unknown constant. 
If the rod is of length l, with the end points at x =* 0 and x =* l, the solu- 
tions of (18-9) must satisfy suitable conditions determined by the mode 
of fixing the ends. If the end x « 0 is clamped, then y{ 0) « ^'(0) * 0; 
if it is simply supported, then y{ 0) = y"(0) = 0; if it is free, then y"(Q) » 
y"'(0) " 0. Similar conditions are imposed at the end x = l. 

For definiteness, we suppose that the ends of the rod are free (a ship 
floating at sea). We then seek a solution of (18-9) such that 

y"( 0) - y"'( 0) « 0 , y''(l) - y"'(0 « 0 . (18-10) 

Since (18-9) is a linear equation, its general solution is the sum of four 
linearly independent solutions 

y(x,\) =* c^i(x,X) + c 2 y 2 (x ) \) + c s y z (x,\) + c A y A (x,\) (18-11) 

where X is the parameter appearing in (18-9) and the c t are arbitrary con- 
stants. On imposing the end conditions (18-10) on (18-11) we get a sys- 
tem of four equations : 

Ciy”( 0 ,\) + £22/2(0, X) + £32/3(0, X) + £42/4(0, X) * 0 , 

£i2/i'(0,X) + c 2 y 2 (0,X) + £32/7(0, X) + c 4 yl'(0, X) = 0, 

+ £ 22 / 2 ( 2 , M + £3 2 / 3 ( 2 , X) + c 4 yl(l } \) = 0, 

£12/7 (2, X) + £22/7(2, X) + £32/7(2, X) + £42/7(2, X) « 0. 

This system of four linear equations in the unknowns c t will have a non- 
trivial solution if, and only if, the determinant D(X) of the coefficients of 
the os is zero. 1 The equation 

X)(X) - 0 (18-12) 

is the characteristic equation , and its solutions are the characteristic values 
of the problem. In general (18-12) is a transcendental equation, and its 
solution poses many vexing problems. 2 Usually it is solved by numerical 

1 See Appendix A. 

1 An instance of a simple transcendental characteristic equation appears in Sec. 10, 
Chap. 6, Eq. (10-16), in which the parameter is denoted by 0. See also Sec. 36, Chap. 1, 
Eq. (36-4), where D(\) ** 0 is an algebraic equation. 



734 


NUMERICAL ANALYSIS 


[CHAP. 9 

methods. Because of the importance of characteristic equations in ana- 
lyzing the behavior of dynamical systems, they have been studied ex- 
tensively and there is a vast literature on the subject of numerical deter- 
mination of characteristic values. 1 

19, Method of Finite Differences. We conclude this chapter with a 
brief description of the most commonly used method for solving boundary- 
value problems in partial differential equations, known as the method of 
finite differences . In this method the differential equation is replaced by 
an approximating difference equation, and the continuous region in which 
the solution is desired by a set of discrete points. This permits one to 
reduce the problem to the solution of a system of algebraic equations, 
which may involve hundreds of unknowns. Ordinarily, some iterative 
technique has to be devised to solve such systems, and high-speed elec- 
tronic computers have been developed largely because of the need for 
coping with problems of this sort. 

The main disadvantage of all numerical techniques is that they give 
numerical values for unknown functions at a set of discrete points instead 
of the analytic expressions defined over the initial region R. Of course, 
when the boundary-value data are determined by measurements at a 
finite set of points of R, the difference-equations methods may be the best 
mode of attack on the problem. Any analytic technique would require 
fitting curves to the discontinuous data. 

We proceed to the outline of the general procedure followed in reducing 
the given analytic boundary-value problem to a problem in difference equa- 
tions. For definiteness let the region R be bounded by a simple closed curve 
C. We seek to determine the function u(r,y) satisfying a given differential 
equation in R. From the definition of partial derivatives it follows that 2 

du u(x + h,y) — u(x,y) 

— — ]j m 

dx h~+ o h 

Also, if the second partial derivatives are continuous one can show that 

d 2 u u(x + h,y ) - 2u(x,y) + u(r - h, y) 

m lim 

dx * h-+o n 

d 2 u u(r + h, y + k) — u(x + h, y) - u(x, y + k) + u(x,y ) 

Jim — , 

dx by o hk 

and so on. 

For small values of h and k the partial derivatives are nearly equal to 

1 For bibliography see Milne, op. cit., and F. B. Hildebrand, "Introduction to Nu- 
merical Analysis/’ McGraw-Hill Book Company, Inc., New York, 1956. 

*See Chap. 3, and Chap. 6, Sec. 21. 



f 


SEC. 19] NUMERICAL INTEGRATION OF DIFFERENTIAL EQUATIONS 735 


the difference quotients appearing in the right-hand members of these 
formulas. If one replaces derivatives in the given differential equation 
by difference quotients, there results a difference equation which is a good 
approximation to the given equation when h and Jc are small. 

Thus, to Laplace’s equation 


V 2 u 


d 2 U d 2 u 

— - -| = 0 , 

dr 2 dy 2 


there corresponds the difference equation 


Ax.eR "h A^u — 0) 

where 

&xxU » [u(x + h,y) - 2u(x ,y) + u(x - h } y)], 

1 

A uv u a -- [v(:r, y + h) - 2 u(x f y) + u(x, y - h)]. 


In a difference equation the values of u(x,y) are related at a set of discrete 
points determined by the choices 
of h and k. Ordinarily these 
points are chosen so that they form 
a square net 1 with specified mesh 
size h. 

The usual procedure is to cover 
the region R by a net consisting of 
two sets of mutually orthogonal 
lines a distance h apart (Fig. 14) 
and mark off a polygonal contour 
C r so that it approximates suffi- 
ciently closely the boundary ('. 

The domain R r in which the solu- 
tion of the difference equation is 
sought is formed by the lattice 
points of the net contained within 
C f . The assigned boundary values 

on C are then transferred in some manner to the lattice points on C\ 
When the lattice points on C do not coincide with points on C, the 
desired values can be got by interpolation. 2 



1 Rectangular, polygonal, and curvilinear nets aie also used. See, for example, 
D. Y. Panov, "Handbook on Numerical Solution of Partial Differential Equations,” 
Moscow, 1051, which contains a good account of the difference-equations techniques. 
See also Appendix to 8. Timoshenko and J. N. (toother's "Theory of Elasticity,” 1051. 

*See, for example, Milne, op> nL, or L. M. Milne-Thomson, "Calculus of Finite 
Differences,” 1933. 


788 


NUMERICAL ANALYSIS 


[CHAP* 9 

One then seeks a solution of the difference equation which satisfies the 
boundary conditions imposed at the lattice points on C'. Usually, this 
leads to a consideration of a system of a large number of algebraic equations 
in many unknowns . 1 

1 Further discussion of difference equations is given in Chap. 6, Secs. 26 and 27, and 
in Chap, 8, Sec. 10. See also chap. 10 by T. J. Higgins in L. E. Grinter (ed.), “Numerical 
Methods of Analysis in Engineering/' 1949. 

The literature on finite-difference methods is extensive. An illustration of the use 
of the method of finite differences in solving a boundary-value problem in Laplace’s 
equation is included in I. S. Sokolnikoff, “Mathematical Theory of Elasticity/' sec. 124, 
McGraw-Hill Book Company, Inc., New York, 1956, which contains further references. 



t 


APPENDIX 




Appendix A. Determinants 

1. The Definition and Properties of Determinants 741 

2. Cramer’s Rule 748 

Appendix B. The Laplace Transform 

1 . Definition of the Laplace Transform 754 

2. Some Uses of the Laplace Transform 756 

3. Discontinuities. The Dirac Distribution 759 

4 Additional Properties of the Transform 762 

5. Steady-state Solutions 765 

6. Integral Equations 767 

Appendix C. Comparison of the Riemann and Lebesgue Integrals 

1. The Riemann Integral 771 

2. Measure 772 

3. The Lebesgue Integral 774 

Appendix D. Table of 4>(x) = J e~ l 12 dt . 


739 




APPENDIX A 


DETERMINANTS 


1. The Definition and Properties of Determinants. A determinant of 
the first order consists of a single element a and has the value a. A de- 
terminant of the second order contains four elements in a 2-by-2 square 
array and has the value 


o,\ a 2 

h b< 2 


= & 1&2 — & 2 & 1 . 


(1-D 


A determinant of third order is similarly defined, in terms of second-order 
determinants : 


0 >\ 0-2 
bx 62 f>3 

Cl c 2 c 3 



b 2 

63 


b 1 


+ a 3 

b 1 

b 2 

= ax 

\ 

c 2 

C 3 1 

— 

Cl 

c z 

Cl 

c 2 

1 


(1-2) 


By analogy, a determinant of order n consists of a square n-by-n array 
of elements (Xi 3 : 



<*12 * 

■ din 

«21 

a 22 * 

• a 2n 

<*nl 

a n2 * 

■ • <*„„ 


to which a numerical value is assigned as follows: Denoting the deter- 
minant by Z), let the elements in the first row be au , and let M\ % be the 
determinant of order n — 1 formed when the first row and ith column of 
I) are deleted. Then, by definition, 

D = a n M n - a 12 M 12 + • • • + (-1 ) 1+n a ln M ln . (1-3) 

The definition is inductive; a determinant of order n is defined in terms 
of those having order n — 1. 

The expansion (1-3) is termed a Laplace development of the determinant 
on elements of the first row. The determinant Mu is called the minor 
of the element an; the signed determinant (~1) 1+ W h is the cofactor of 
Uxt, More generally, the determinant Mi 3 formed when the fth row and 

741 



APPENDIX 


742 


[app. a 


jth column are deleted is the minor of the element a# in this row and 
column. The signed determinant 

(-1)^% (1-4) 


is the cofactor of a*/. It is a fundamental theorem that a determinant may 
be evaluated by a Laplace development on any row or column; in other words, 


an 

Oi2 * 

* «ln 



02 1 

&22 ' 

* 02n 

n 

~ OijAij = 

Z <hjA 

7-1 

«nl 

0*2 • 

’ * o nn 




The proof may be given by induction directly or may be based on the following 
considerations, which are also established by induction. The expansion of an nth-order 
determinant is a sum of the n! terms ( — l)*a* . . .a* n „, where k\, . . M k n are the 

numbers 1,2, . . . , n in some order. The integer k is defined as the number of inversions 
of order of the subscripts k\ t k 2 y . . k n from the normal order 1, 2, . . n where a par- 
ticular arrangement is said to have k inversions of order if it is necessary to make k suc- 
cessive interchanges of adjacent elements in order to make the arrangement assume the 
normal order. There are n * terms, since there are n! permutations of the n first sub- 
scripts, and each term contains as a factor one, and only one, element from each row 
and one, and only one, element from each column. 

For example, consider the third-order determinant 


011 

012 

018 

021 

022 

028 

031 

032 

033 


The six terms of the expansion are, apart from sign, 

011022033, fl U a 32023, <*21012033, 021033013, 031012028, Og 1022013- 

The first term, in which the first subscripts have* the normal order, Is called the diagonal 
term, and its sign is positive. In the second term the arrangement 132 requires the 
interchange of 2 and 3 to make it assume the normal order; therefore k *» 1, and the 
term has a negative sign. Similarly, the ttiird term has a negative sign. The fourth 
term has a positive sign, for the arrangement 231 requires the interchange of 3 and 1 
followed by the interchange of 2 and 1 to assume the normal order. Similarly, the 
fifth term has a positive sign. In the sixth term, it is necessary to make three inter- 
changes (3 and 2, 3 and 1, and 2 and 1) in order to arrive at the normal order; hence, 
this term will have a negative sign. It follows that 

D « 011022033 — 011032023 — 021012033 “f 021032013 4“ 0*1012028 — 081022018. 

The main result of this discussion is that a determinant is the sum of 
all the nl products which can be formed by taking exactly one element from 
each row and each column and multiplying by 1 or —1 according to a 
definite rule. 



SEC. 1] DETERMINANTS 

Example 1. By a development on the second column, evaluate 

10—12 
I a 
D 


743 


6 0 
4 7 
2 0 


The ( — l)*** rule for determining sign means that the sign of the minors alternates as 
we proceed from one element to an adjacent one in the same row or column, and the 
sign starts with 4- in the upper left-hand corner. Thus, 

D m O(-Afia) + O(Afa) + 7(-Af w ) + 0 (Af«). 

Crossing out the row and column containing 7 gives the determinant M&, whence 


D - — 7 JV7 32 - (—7) 


1 -1 2 
6 4 3 

2 2 3 

( — 7)f(I)(G) — ( — 1)(12) 4-2(4)] 


( — 7)[(l)j 


4 3 
2 3 


(-D 


6 3 
2 3 


+ ( 2 ) 


6 4 
2 2 


-182. 


D 


Example 2. The following determinant is said to be in diagonal form. Show that its 
value is abed no matter what elements are put in place of the *s: 

a * * 

0 6 * 

0 0 c 
0 0 0 d I 

Successive l^aplace developments on first columns give 

6 

i 

D - a 


0 c * 
0 0 d 


ab 


c w 

0 d 


abed. 


Evidently, a similar result is true in general. 

Example 3 If the elements are differentiable functions of t , show that the derivative 
of the determinant (1-2) is 


/ 

/ 

02 

t 

a 8 


ai ag 03 


Ol 

02 

as 

h 

h 

63 

+ 

K K K 

4- 

bi 

1)2 

6s 

Cl 

C2 

cz 1 

1 1 

Cl C2 C 3 | 


c i 

c 2 

*3 


A typical term in the expansion is ±a t bjCi. Differentiating gives 
±(a t bjc k ) r » dr a'ibjCk ± a x b’ } c k db a t 6 ; c*, 

and the sum on i, j, k of these three types of terms yields the expanded form of the 
three determinants. A corresponding result for determinants of order n is proved in the 
same way. 

The fundamental theorem (1-5) leads to some important properties of 
determinants that are now enumerated. 

1. If each element in a row or column of a determinant is zero , the deter - 
min ant is zero . 



744 APPENDIX [app. a 

2, If each element in a row or column is multiplied by m, the determinant 
is multiplied by m. 

3. If each element of a row or column is a sum of two terms , the determinant 
equals the sum of the two corresponding determinants; for example , 


ai 

h 

Cl 


ax 

h 

Cl 


a t 

bx 

Cl 

a + ot 

b + fi c + y 

- 

a 

b 

C 

+ 

a 

0 

y 

<h 

b 2 

c 2 


<h 

b 2 

c 2 


a 2 

b 2 

C2 


These three results become obvious when we make a Laplace develop- 
ment on the row or column in question. In (1-6), for example, let A , B, C 
be the cofactors of the elements in the second row. The determinant is 

(a + a)A + (b + 0)B + (c + y)C , 

which equals {a A + bB + cC) -f (c*A + $B + y C). This, in turn, is the 
sum of the expansions of the two determinants on the right of (1-6). 
The proof for n-by-n determinants is very similar and should be supplied 
by the reader. 

4. If two rows or two columns are proportional , the determinant is zero. 

5. If two rows or two columns are interchanged , the determinant changes 
sign. 

6 . If rows and columns are interchanged , the determinant is unaltered. 

The properties 4, 5, and 6 are easily verified for 2-by-2 determinants, 
then proved in general by mathematical induction. To obtain item 6, 
for example, expand the original determinant on elements of the first row 
and the new one on elements of the first column. The theorem for order 
n then follows from the theorem for order n — 1. As an illustration we 
have 



b 1 

Cl 

a 2 

b* 

C2 

03 

h 

c 3 


b 2 C 2 

&3 c 3 


— 


b i 
bs 


c 1 


+ «3 


b i c x 
&2 c 2 


(1-7) 


which coincides with the expansion (1-2) when we interchange rows and 
columns of the second-order determinants on the right-hand side of (1-7). 

7. The value of a determinant is unaltered if a multiple of one row (or 
column) is added to another. 

8. If the cofactors for one row (or column) are combined with the elements 
of another y as in (1-8), the resulting sum is zero: 

n n 

4 ^ ftijAik 0 , ajiAfa “ 0 , k 3r^ j. 

ifml i- 1 


( 1 - 8 ) 



SBC. 1] DETERMINANTS 746 

These results follow from those already established. To illustrate the 
proof of item 7, we have 

oi bi + mci ci a x bi c x a x mci c x 

a 2 + rnc 2 c 2 ~ a 2 b 2 c 2 + a% mc 2 c 2 (1-9) 

a z bz 4“ fnc z c 3 #3 ^3 c 3 03 meg C 3 

by 3; and the second determinant on the right of (1-9) is zero by 4. The 
reader should extend the proof to n-by-n determinants. 

The result 8 follows from 4. Thus, the first expression (1-8) is the 
expansion of the determinant which arises when the row 

a lk) a 2k) • • Gnfc 

is replaced by a^, a 2 j, . . . , a n *, and hence it is the expansion of a de- 
terminant with two equal rows. 

9. If two determinants A and B of order n are given and a new determinant 
C is formed, the element in the ith row and jth column of which is obtained 
by multiplying each element in the ith row of A by the corresponding dement 
in the jth row of B and adding the products thus formed, then C = AB. 

Thus, if the elements of A and B are denoted by a tJ and respectively, 
then the element c XJ in the ith row and jth column of the product deter- 
minant C is 

c t} — a 4 i&ji + a l2 bj 2 H b a xn b jn . (1-10) 

The validity of rule 9 for determinants of order n follows from considera- 
tions entirely similar to those we give next for the case when n = 2. 

If the determinants A and B are of second order, formula (1-10) states 
that their product C is 

Ull&ll 4" &12&12 G11&21 + tt 12&22 
a 2lhll + a 22^12 #21^21 4“ U22&22 

Since the elements in (1-11) are binomials, we can write C by using prop- 
erty 3 as the sum of four determinants: 

an&n 011&21 ^12^22 

021&11 021&21 a 21^Xl a 22^22 

012&12 a ll&21 G12&12 a l$22 

a 22^12 a 2l^2l °22&12 <*22^22 

On factoring out the elements an and a 2 i in the first determinant, we 
obtain a determinant with two like columns, and hence its value is zero. 
Similar remarks apply to the fourth determinant. The second deter- 




( 1 - 11 ) 




APPENDIX 


746 


{app. a 


minant, on factoring out fen and fe 2 2 , yields fenfe 2 2 A, while the third haa 
the value —5^21*4. Thus 


C *= A (fej ife 2 2 — fei2fe2i) 


- AB . 


Similar, but much more laborious, calculations can be carried out for deter- 
minants of higher order to establish the validity of rule 9. 

Since the value of a determinant is unchanged when its rows and columns 
are interchanged, there are four ways in which the determinant C may 
be written. Thus, if we interchange the rows and columns of B, the ele- 
ments on in C will be given by (1-10) in which the subscripts on the fes 
are interchanged. 1 


Example 4. Without expanding show that 


1 x x x\ 

l Xz x\ 

1 xz x\ 


(*i - x 2 )(x 3 - x 2 )(xi - Xz). 


The determinant is a polynomial in xi, and it vanishes when x\ » x 2 , since the first two 
rows are then proportional. Hence it is divisible by 2*1 — X 2 . Similarly, it is divisible 
by xz — xj and xi — xz. It therefore equals 

E(x 1 - X 2 )(X3 ~ X 2 )(Xi ~ Xz) 


for some polynomial E. Since the determinant is of degree 3 in Xi, X2, X3, we must have 
E » const, and comparing coefficients of x 2 x \ shows that E — 1. 

Example 5. Write the product of the determinants 



1 

2 

1 



-1 

4 2 

A »= 

3 

0 

1 

and 

B » 

2 

-1 3 


' 0 

2 

1 1 



0 

2 -1 


as a single determinant of third order. 
Using rule 9 we find 


AB 


-1 +8 + 2 2-2 + 3 0 + 4-1 
-3+0+2 6-0+3 0+0-1 
0+8+2 0-2+3 0+4-1 

9 3 3 

-19-1* 

10 1 3 


To check the result, we find on expanding the determinants A } B t and AB that A * —2, 
B « 21, and AB * —42. 

Example 6. Show that a trigonometric polynomial 

y « ai sin x + a% sin 2x + sin 3x 
1 Cf, Eq. (16-3), Chap. 4. 


(M2) 



SEC, 1] DETERMINANTS 

passing through three assigned points {z % ,y t ) is given, in general, by 


y sin a: sin 2x sin 3x 
yi sin xi sin 2xi sin 3xi 
2/2 sin sin 2xj sin 3x2 
y% sin x$ sin 2 x 3 sin 3xa 


0. 


Expanding on the first row gives 


747 


c\y -f* ci sin x -f C 3 sin 2x 4* c% sin 3x — 0 

where c* are the appropriate cofactors. Hence y has the form (1-12), if c\ 9 * 0. More- 
over, when x « x, and y « y t the equation is true, since two rows of the determinant 
are then equal. 


PROBLEMS 


1. By a Laplace development on the first row, evaluate 


1 2 3 


— 1 2 2 


1 0 0 


1 2 3 

3 1 2 

» 

-3 6 6 

> 

0 0 1 

, 

4 5 6 

2 3 1 


5 7 9 


0 10 


7 8 9 


2. Evaluate the determinants in Prob. 1 by a Laplace development on (a) the first 
column, (6) the second row. 

3 . Evaluate this determinant by development on 


(а) The first column 

(б) The second row 
(c) The first row 

( 1 d ) The third column 


1 -1 

0 1 

0 0 

1 0 


1 -1 

-1 1 

1 -1 

0 1 


4 . Show that 


zi 1 
x 2 1 


Xj 

V\ 

1 


V 2 

1 

^8 

ys 

1 


represent, respectively, the (signed) length of the segment (xi,X 2 ) and the area of the 
triangle with vertices (x,,y»). 

6. Evaluate, using some of the properties 1 to 7: 


X 1 1 


y 4- z x x 


0 -a -b 

1 X 1 

1 

y x + * y 

} 

0 

1 

r> 

1 1 X 


z z x 4- y 


be 0 


Hint' In the last determinant, interchange rows and columns. 

6. Write out as determinants of third order the product of the first determinant in 
Prob. 1 by the second and third determinants. 

7. Using determinants, find a, b, c if y ■» a -f b cos x + c cos 2x passes through (0,0), 
(r/2,1), (*•, -2). 

8. (a) Find a cubic containing the points (0,1), (1,-1), (3,4), (4,0). Hint: Consider 
a determinant with top row y, 1, x, x 2 , x 3 . (6) Write the equation of a polynomial of 
degree n whose graph contains n + 1 assigned points (x*,y t ). 



748 


APPENDIX 


[APP. A 


8. Cramer’s Rule. Consider the set of simultaneous equations 
a\X + b\y + C\t *= 

<hx + b 2 y + c 2 z ® d 2t 
dzx + b$y + c 3 z » da. 

Now, by 2 and 7 of the preceding section, 


(2-1) 



ai 

61 

Cl 


aix 

h 

Cl 


diX + b x y + c x z 

b 1 

Cl 

X 

a 2 

bi 

c 2 

- 

a 2 x 

62 

C 2 


a^x + b%y -|- dz 

62 

c 2 


a 3 

b 3 

c* 



&3 

cz 


a 3 x + b 3 y + c 3 z 

b 3 

Cz 


Hence if x satisfies (2-1), it is necessary that 



ai 

h 

Cl 


d 1 

b 1 

Cl 

X 

a 2 

^2 

C2 


d 2 

h 

C 2 


^3 

bs 

C 3 


d 3 

b 3 

c 3 


(2-2) 


The determinant on the left of ( 2 - 2 ) is termed the coefficient determinant 
of the system ( 2 - 1 ); we denote it by D. Equation ( 2 - 2 ) and the cor- 
responding relations for y and z may then be written 


(2-3) 


If D 5^ 0 , we may divide by D to express x, y ) and 2 as quotients of two 
determinants. 


xD — 

d\ bi Ci 

d 2 b 2 c 2 

■ yD = 

&i d\ Ci 

a 2 d 2 c 2 

II 

d\ b\ d\ 
cc 2 b 2 d 2 


d$ b% c 3 


\ <^3 d 3 C3 


&3 d 3 


To show that these values of x t y, and z actually satisfy the system (2-1), substitute 
into (2-1) and multiply through by D. The equations become 



d\ 

bx 

°i ; 


ai 

di 

C\ 


| 

bi 

di 


1 «i 

b 1 

Cl 

Oifcj 

<k 

b 2 

02 

+ 6*j 

02 

d 2 

C2 

4* c k 


b 2 

di 

! - * 

02 

bi 

C2 


di 

b, 

Cl I 


a% 

di 

Cl 


ai 

bi 

di 


0 * 

b. 

Cl 


with k «■ 1 , 2 , or 3 respectively. Now, the determinant 

au bk Ck dk 
ai bi a d\ 

h <% d% 

<H b§ ci dd 

jft *ero because twp rows are equal, and it yields the desired relation when expanded on 
elements of the first row (use Theorem 5 of the preceding section). 



BBC, 2] DETERMINANTS 749 

The foregoing method applies to n equations in n unknowns, and yields 
Cramer’s Rule: Let 1 


Oll^l + 012^1 + • * * + d\ nX n = &l, 

d 2 \Xi + a 2 2^2 d 1- d 2n x n = k 2l 

(2-4) 


Gnl%l + d n 2 X 2 H f" d nn X n *= & n 

be a system of n equations in the n unknowns x x such that the coefficient de- 
terminant D is not zero. The system (2-4) has a unique solution x t = D,/Z), 
where D t * is the determinant formed by replacing the elements a Ul a 2l , . . . y a n % 
of the ith column of D by k x , k 2y . . k n respectively. 

Consider the homogeneous system which arises from (2-4) when the 
right-hand members are replaced by zero. This system obviously has a 
solution X\ = x 2 = * • • = x n = 0, the trivial solution. If the coefficient 
determinant is not zero, the solution is unique by Cramer's rule. Hence 
a homogeneous system can have a nontrivial solution only if the coefficient 
determinant is zero. One can prove, conversely, that there is always a 
nontrivial solution of the homogeneous equations if the determinant is 
zero. 

The rectangular array 

( ai bi Ci di\ 

a 2 b 2 c 2 d 2 I (2-5) 

a 3 h c 3 d-J 


is termed the augmented matrix of the system (2-1). By striking out one 
or another column of the matrix (2-5), we are led to the square arrays 



'dl 

b , 

Cl \ 

a 2 


7 


&3 

cj 


Since these arrays are square, they have corresponding determinants. 
Now (2-3) shows that all these determinants must be zero if D * 0 and if 
the system (2-1) actually has a solution. In other words, if D * 0 but 
a third-order determinant formed from (2-5) is not zero, then the system 
(2-1) is inconsistent. 

The foregoing results are included in a general theory of linear systems, 
which is now discussed. An ra-by-n matrix is a system of mn quantities 
a x j, called elements, arranged in m rows and n columns. The array is eus- 

1 A compact derivation of this rule is given in Sec. 15, Chap. 4. 



[app. a 


750 APPENDIX 

tomarily enclosed in parentheses, thus; 



Mil 

a 12 * 

’ ’ a ln \ 

A SE 

f <*21 

1 * * 

<*22 * 

* * «2n 


\<*ml 

a m2 * 

’ ’ <*mn/ 


If m «= n, then A is the coefficient matrix of the system (2-4) ; the augmented 
matrix is obtained by adjoining a column with elements (in order) k\, 
A' 2 , . . . , k n . If the matrix is square, one can form the determinant of the 
matriXj a determinant whose elements have the same arrangement as 
those of the matrix. From any matrix, smaller matrices can be formed 
by striking out some of the rows and columns. Certain of these smaller 
matrices are square, and their determinants are called determinants of the 
matrix. A matrix A is said to he of rank r if there is at least one r-rowed 
determinant of A that is not zero, whereas all determinants of A of order 
higher than r are zero or nonexistent. (The latter alternative arises if r 
equals the smaller of the two numbers m and ft.) The rank is zero if all 
elements are zero. With these preliminaries we can state the following 
Fundamental Theorem: Suppose we are given a set of m linear equations 
in n unknowns. Let the rank of the coefficimt matrix he r, and let the rank of 
the augmented matrix he r' . If r ' > r, the equations have no solution . If 
r' ss r = n, there is one, and only one , solution. If r' = r < n, we may give 
arbitrary values to n — r of the unknowns and express the others in terms of 
these . 

The proof is too long for inclusion here. Important special cases were 
established, however, by the proof of Cramer’s rule and by the discussion 
of (2-5). Further discussion of matrices is given in Chap. 4. 

The r unknowns which are expressed in terms of the others must be asso- 
ciated with some nonvanishing determinant of order r. 


Example 1. By Cramer’s rule, find x and y, given 

3x 4" y 4* 2s; «* 3, 

2x — 3y — z ** 3, (2*6) 

x ■+* 2y + z » 4. 

The coefficient determinant D is found to be 8, so that 



3 

1 

2 



3 

3 2 

Sx ~ 

-3 

—3 

-1 

-8, 

Sy « 

2 

—3 -1 


4 

2 

I 



1 

4 1 


Thus, x m 1, y m 2. If z is desired, one can find it from the third equation (2-6): 
z » 4 — x — 2j/-4 — 1— 4 — — 1. 



BBC. 2 ] DETERMINANTS 

Example 2, For what values of X do the equations 


751 



a 2 * + b 2 y ~b C 2 = 0. 

Let (x y y) be the point at which the three lines meet. With this particular choice of 
x and y, the three equations (2-12) are satisfied simultaneously. Now, these equations 
may be regarded as simultaneous equations in three unknowns x , y t and 1, one of which 
(namely, 1) is not zero Hence the coefficient determinant vanishes: 

a b c 
°i b i c\ ** 0. 
a 2 b% C2 



752 


APPENDIX 


[AFP, A 

(The condition is also sufficient if no two of the three lines are parallel The reader 
should observe the duality between points and lines which is illustrated by this and 
the following example.) 

Example 5. Find a necessary and sufficient condition that the three points (x,y), 
(xj^s) lie on a line. 

If the equation of the line is 

ax 4* by 4* c *» 0 (2-121) 

we have, besides (2-13), 

oxi + byi 4* c » 0, 
axt + f >|/2 4- c « 0. 

These equations may be regarded as a system in the unknowns a, b t c, which cannot 
all vanish if (2-13) represents a line. Hence the coefficient determinant must vanish: 

x y 1 

xi y x 1 « 0. (2-14) 

x% yi 1 

Conversely, (2-14) ensures that the system has a nontrivial solution a, b , c. Compare 
Prob, 4, Sec. 1 . 

Example 6. Show that the following equations are consistent if, and only if, k « 9: 

2x 4* 3y » 1, 

* - 2y - 4, (2-15) 

4 x ~~ y ** k. 

The coefficient matrix has rank 2, and hence the equations are consistent if, and only if, 
the augmented matrix also has rank 2. This entails 

2 3 1 

1 -2 4 - 0, 

4 -1 k 1 

which yields 1(7) — 4( — 14) + k(— 7) — 0 or k — 9. The same result is found if we 
regard (2-15) as a system in the three unknowns x, y, k and solve by Cramer’s rule. 
The reader should obtain the result of Examples 3 and 4 by considering the augmented 
matrix, as in the present example. 

PROBLEMS 

X. Solve, by Cramer’s rule, the systems: 

(a) x 4“ 2y ~h 3s «■ 3, (b) 2x 4- y 4* 3z ** 2, 

2x — y -f z » 6, 3x - 2y — 2z * 1, 

3x4 - y ~ z ** 4; x — y 4 - * - — 1; 

(c) x 4- 2y « 1, (d) 2x + y 4- 3z 4- w * -2, 

2x - y — 2z ® 3, 5x 4- 3y — x — w * 1, 

—x 4* V + 8* ** 2; x — 2y 4- 4x 4- 3w? » 4, 

3x - y 4- 2 * 2. 



SBC* 2] DETERMINANTS 

2. Obtain nonxero solutions when they exist. 


753 


(a) x 4* 3y — 2* * 0, (b) x — 2y » 0, 

2a? — y 4* * ** 0; 3x 4- 3/ •* 

2x — y *» 0; 

(e) 3x — 2y + z - 0, (d) 2x — 4y 4- 3z ** 0, 

x 4- 2y — 2z *® 0, x + 2y - 2z - 0, 

2x - ?/ 4- 2z « 0; 3x - 2y + * - 0; 

(e) 4x - + z * 0, (/) x + 2y + 2z - 0, 

2x — y -f 3z « 0, 3x - y 4- z * 0, 

2x — y — 2z *» 0, 2x 4~ 3y + 2z « 0, 

fix - 4* 4z « 0; x 4- 4y ~ 2z = 0. 

3. Investigate the following systems and find solutions whenever the systems are 
consistent: 

(a) x - 2j/ - 3, (b) 2* 4- y - * - 1, 

2x 4- y «* If x — 2y 4~ z *■ 3, 

3x — y » 4; 4x - 3^ 4- « * 5; 

(c) 3x 4- » 4, (d) 2x — y 4- 3z «* 4, 

x - * 1, x 4- V - 3z » -1, 

2x 4- by ** — 1 ; 5x - 2 / 4* 3z *» 7. 

4 . (a) Give a necessary and sufficient condition that four points in space be coplanar. 
(6) Give a necessary and sufficient condition that four planes be concurrent. 

5. As in Example 5, find a necessary and sufficient condition that four points lie on a 
circle, 

6. Give a relation which the coefficients must satisfy if 

ox 4 5 6 7 8 9 10 4- &x 2 4- cx 4- d « 0, 
ax 2 4- /3x 4- 7 w 0 

have a common root. 

7. Give a condition on the coefficients of a general cubic f(x) if it has a double root. 

and /'(x) have a root in common. 

8. The system ax 4- by « c, ax 4- fly ** 7 represents two lines which may intersect 
at one point, may be parallel, or may coincide. Discuss the system geometrically, and 
thus obtain all the relevant results involving rank. Hint : Begin by showing that the 
lines are parallel if, and only if, the coefficient determinant is zero. 

9. An equation ax 4“ by 4~ <% “ d represents a plane, and two planes are parallel if, 
and only if, corresponding coefficients a, 6, c are proportional: 

a ** kai, b » kb i, c *» kc\. 

(You may assume these geometric facts.) As in Prob. 8, give a complete geometric 
discussion of the behavior of two equations in three unknowns. 

10. As in Prob. 9, discuss the general system of three equations in three unknowns. 



APPENDIX B 


THE LAPLACE TRANSFORM 


The use of Laplace tranforms for solving ordinary differential equations 
has its origin in a symbolic method developed by the English engineer 
Oliver Heaviside. It enables one to solve many problems without going 
to the trouble of finding the general solution and then evaluating the arbi- 
trary constants. The procedure can be extended to systems of equations, 
to partial differential equations, and to integral equations, and it often 
yields results more readily than other techniques. 

1. Definition of the Laplace Transform, The function F(p) given by 

F(p) « f* f(x)e~ px dx « L(/) (1-1) 

-'o 

is called the Laplace transform of f(x), and the operator L that transforms 
/ into F is called the Laplace transform operator. The operator L is linear ; 
that is, 

L(f+g) = L(f) + L(g), (1-2) 

Ucf) - cL(f), (1-3) 

where c is any constant. Indeed, the definition of L shows that (1-2) is 
equivalent to 

[ tf( x ) + g(x)]e~~ pz dx « f f(x)e~ px dx + f g(x)e~ pz dx 
Jo Jo Jo 

and this is a familiar property of integrals. The proof of (1-3) is similar. 

To illustrate the calculation of a Laplace transform let f(x) = e az , where 
a is constant. The transform is 

/ e ax e-* z dx « / <r<P~ a >* dx = 

Jo Jo — (p — a) 

provided p > a. When p < a, the integral diverges. 

This example enables us to investigate the convergence of (1-1) for a 
general function /(z), provided 


(1-4) 


764 



sue. 1] 


THE LAPLACE TRANSFORM 


755 


fix) is piecewise continuous 1 on every finite interval 

and (1-5) 

1/0*0 1 < Me ax for some choice of the constants M and a. 

Under these conditions the integral converges for p > a , just as in the 
foregoing example. In fact, 

f* \f(x)\e" px dx < f* M c ax e~ px dx < M f* dx. 

Since the latter integral has the finite value (1-4), the integral on the 
left remains bounded as l —► «. This establishes not only the conver- 
gence but the absolute convergence of the integral defining L (/). The 
convergence is uniform if p > ao > a ) where a# is fixed, and hence the 
operations we shall carry out later are justified. 

The integral on the right of the foregoing inequality tends to zero as 
p —* oo. This allows that 

lim F{p) = 0 (1*6) 

p— f 00 


for all functions F = L (J) such that / satisfies (1-5). It is found, more 
generally, that F(p) — > 0 if L(/) converges for any finite value p = p 0f 
even when (1-5) does not hold. Hence, if lim F(p) 5 * 0 as p 00 , then 
F{p) cannot be the Laplace transform of any function /(r). 

Example 1. Let f(x) ** x b . The change of variable t = px yields 


P 


x b e pl dx 




According to Chap. 2, Sec, 14, the latter integral is convergent for b > —1 and repre- 
sents the generalized factorial 6!. Hence 

L(x 6 ) - for 6 > -1. (1-7) 

When b is negative, x h is infinite at x * 0 and (1-5) does not hold. 

By comparing the integral for L(/) with that for L (Mx b ) near the origin, one finds 
that (1-5) is really needed only for x > 1, provided f(x) is piecewise continuous for 
x > 0 and satisfies the additional condition 


| f(x) | < Mx b on 0 < x < 1 for some constant 6 > — 1. 

Whenever we take a Laplace transform L (/) in the sequel, it is understood that p > a 
and that/ satisfies (1-5) or the more refined condition just described. On the other hand 
it is not required that f(x) be real For example, (1-4) holds when a is complex provided 

p > Re (a). 

1 See Chap. 2, Sec. 25. The following discussion uses a comparison test for integrals, 
which can be verified in the same way as the corresponding feet for series. Cf. Chap. 2, 
Sec. 4, Theorem I, and Chap. 2, Sec. 6, Theorem I. 



766 APPENDIX 

Example 2. The choice a «■ ib in (1-4) yields 


[APP. 8 


L(e**) _ L(oos bx + i sin for) — • 


p — ib 

Upon equating real and imaginary parte with due regard to (1-2) we get 


P 

t>(coa bx) ■* —x ~zi L(sin bx) 

P + b* 

for all real 6. Differentiation with respect to b gives 


L(x cos bx) 


p 2 -b 2 

(p 2 +bY 


L(x sin bx) 


b 

P ! + 6* 
2 bp 


(P* + 6*) J 


( 1 - 8 ) 


(1-9) 


Proceeding in this fashion one can construct a table of transforms, such as Table 2 
given at the end of this appendix. Indeed, we have already derived entries la, 2a, 26, 
36, and 4a of Table 2; and entry 3a can be obtained from (1-8) and (1-9), since L is linear. 


2. Some Uses of the Laplace Transform. If L[f(x)] = F(p), integration 
by parts leads to 

L[f(*)] = pF(p) -m (2-1) 

provided the hypothesis (1-5) applies to f'(x) as well as tof(x). That is, 

00 

f e~ p *f'(x) dx = e~ px f(x) + f pe~ px f(x) dx. (2-2) 

A) o ■'0 


For sufficiently large p Eq. (1-5) shows that e~ pz f(x) — > 0 as x oo, and 
the desired result follows. 

The choice /(x) — y in (2-1) gives 

My') = pL(y) - y( 0 ) ( 2 - 3 ) 

and the choice f(x) = y' gives 

My") * pl^y') - 2/'(0) ** p[pL(.v) — p( 0 )] - y\ 0 ) 

in view of (2-3). Hence 

My") = p 2 L(^) - py( 0) - y'(0). (2-4) 

The transform of the higher derivatives can be obtained similarly. For 
instance, 

My'") « v*My) - p 2 2/(0) - py'( 0 ) - y"{ o). (2-5) 

These relations enable us to solve differential equations with constant 
coefficients. 

As an illustration consider the problem 

y" + y * /( 0 » 2/(0) “ y'{Q) - 0 , ( 2 - 6 ) 

which describes the response of a resonant circuit to an input /(/) . To make 
the problem definite let/(2) * 0 for t < 0, but f(t) « 1 for t > 0. (A switch 



SEC. % THE LAPLACE TRANSFORM 757 

to a constant-voltage source is closed at time t « 0 and remains closed there- 
after.) The transform of (2-6) gives 

p 2 Uy) + L(y) « L(/) - p~ l 


when we use (2-4) and the entry la of Table 2 with a 

Uy), 

Hy) 


l 


vir + 1 ) 


0. Solving for 


It can be shown that a continuous function y is determined on (0,«>) 
as soon as its transform L (y) is known. Hence, the foregoing equation 
contains the solution implicitly. To find the solution explicitly we use 
partial fractions; thus 

v 1 V 

Uy) T7™ 

V P + 1 


The entries la and 26 of Table 2 give the desired answer 

y = 1 — cos i for t > 0, y = 0 for t < 0. (2-7) 

It is an especial merit of the Laplace transform that the initial conditions 
are satisfied automatically. In the foregoing illustration we did not find 
the general solution and then determine the constants so that y( 0) = y'{0) 
— 0. Nevertheless, the expression (2-7) satisfies these conditions, as the 
reader can verify. 

To illustrate further the introduction of initial conditions we shall solve 


y'" -y' ~ sin x (2-8) 

subject to 

1/(0) = 2, y\ 0) = 0, 2/"(0) * 1. (2-9) 

The Laplace transform of (2-8) yields 

p z L(y) - 2p 2 - l ~~ \pL(y) - 2] = L(sin x) » (p 2 + l)~ x 
when we use (2-5), (2-3), and entry 2a of Table 2, Solving for L (y) t 


Uv) 

By oartial fractions 

Uy) - 


2p 2 - 1 


+ 


(p 2 + i)(p 3 - p) 

3 V 


+ 


4 (p + 1) ' 4(p - 1) ' 2(p 2 + 1) 
and entries la and 26 of Table 2 give 

V “ He~ x + %e* + H coax. 



768 APPENDIX [app. b 

The Laplace transform can also be used to solve systems of differential 
equations. As an illustration, let it be required to find y if 


y' + 2z' -f y — z - 25, 

2y' + 2 = 25 e 1 , 

with the initial conditions 

y(0) = 0, *(0) = 25. 

The transform of (2-10) leads to 

P L(y) + 2[pL(z) - 251 + L(y) - L (z) = 


25 


2pHv) + L (z) = 

which simplifies to 

(p + l)L(p) + (2 p - l)L(z) 


25 


p - 1 

25 (2p + 1) 


Solving for L (y), we get 

Uy) 


p 

2pL(p) + L(z) = 25(p - l) -1 . 
25 


4 p(p - l) 2 (p + 


( 2 - 10 ) 

(2-11) 


_ 25 9 5 16 

p p - i + (p - i) 2 p + H 
According to' entries la and 15 of Table 2, 

y = 25 - 9r* + bjcc x - Hie-*' 4 . 


It should be noted that this method enables us to find y without finding 
z. Also no extraneous roots are introduced, and the initial conditions are 
satisfied automatically. 

PROBLEMS 


1. If y satisfies y " — 3 y r *f 2y — 4, y(0) « 2, y'(0) ■» 3, show that 


Uv) 


2 p 2 - 3p + 4 
p(p ~ 1)(P ~ 2) 


Deduce that y « 2 — 3e x -f 31? 2 *. 

2 . Solve by means of the Laplace transform 


V " + 4y * sin z, y(0) - 1, y'( 0) - 0. 

8. Find L(y), and solve 


8'" + *"-«*+* + 1, 2/(0) - |/'(0) - y"(0) - 0. 



759 


tSEC, 3] THE LAPLACE TRANSFORM 

4 . Find L(«) in Eqs. (2-10) and (2-1 1) of the text, and deduce that 
2 « 336* - lOxe* - 

6. Solve by means of the Laplace transform and check by substituting into the given 
system: 

V' + 3y -f 2 ' 4* 2z - e~ 2z , y( 0) • 0, 

2y' -F 2y 4- *' + 2 • 1, *(0) * 0. 

6. Find y, given that 

y' -f z' « z' -f v) f » w' + y' *=* y, y(0) ** 2(0) = tt>(0) « 1. 


7. If /'(j) satisfies (1-5), show that fix) satisfies a condition of the same type, though 

perhaps with a different value of a. Hint: f(x) « / f'(t) dt -f /( 0). 

J 0 


3. Discontinuities. The Dirac Distribution. Closing a switch in an 
electrical circuit introduces a discontinuity in the corresponding input 
function [cf. the discussion of (2-6)]. A disconti- 
nuity may also be produced by a sudden impulse 
in a mechanical system. The Laplace transform 
is a most effective means of dealing with such situ- 
ations, because the transform of many discontinu- 
ous functions is just as simple as L(e*) or L(sin x). 

Tn this section we shall consider the response 
of a system to an impulse function which acts 
over a very short time interval but produces a 
large effect. The physical situation is typified by 
a lightning stroke on a transmission line or by a 
hammer blow on a mechanical system. 

To formulate the idea of an impulse, let a be 
a small positive constant and let S a (x) be the function illustrated in 
Fig. 1. That is, 

6 a (x) = a~~ l for 0 < x < a 
and S a (x) =* 0 elsewhere. The Laplace transform is 



L[5 a (x)] * \ a a~ l e~ px dx « (pa)~ l ( 1 - eT pa ). 
J 0 


By the Taylor series for e pa 

“ 1 — M(p a ) 4 — 1 

as a — ► 0 . It is customary to introduce an expression 6(x) which is thought 
to be the limit of 6 a (x) as a — ► 0 and to say that 

m*)] - 1 . (3-i) 

We call 5(x) the Dirac distribution or the unit impulse , and we take (3-1) 
as the basic defining property. The legitimacy of this procedure requires 



APPENBIX 


760 


[app. b 


further discussion, which will be given presently. First, however, the use 
of S(x) will be illustrated by an example. 

The displacement y of a weight suspended by a spring with stiffness 1 
is determined by the system 

y” + v « f(t), y{ 0) - v'( 0) « o, 

where f(t) *» force function 
t « time 
' * d/dt. 

To determine the response to a unit impulse at t « 0 we replace f(t) by 
5(0; thus 

y" + y « *(Q- 

The Laplace transform yields 

p 2 1(v) + Uy) = L[3(0] - l 
when we use (3-1). Hence L {y) « 1/(1 + p 2 ), or 

y -» sin t, t > 0. 

The initial conditions require y = 0 for t < 0, and the graph has the ap- 
pearance illustrated in the accompanying Fig. 2. 

The function y is continuous, 
but it is not differentiable at 
t - 0. Thus, the initial condition 
y'(0) — 0 is not satisfied. Indeed, 
y'(l) — ► 0 as t 0 through nega- 
tive values, but y'(i) = cos t 1 
as t — ► 0 through positive val- 
ues. The unit impulse produces 
a jump, of magnitude 1, in y'{t). 

To investigate the meaning of the foregoing result, we solve 
y" + y ~ 8 a (t), 2/(0) ** y'(0) - 0 

and then let a —► 0. The general solution is 

y « co sin t + ci cos i, t < 0, 
y « C 2 sin t -b c 3 cos t *b a - * 1 , 0 < t < a, 

2/ *» C4 sin f 4* C5 cos t, t > a. 

By the initial conditions, 

Co * Cl * Cj * 0, C8 «* 

a 

To determine c* and c* we require that y and y’ be continuous at t «* a. This gives 
—a"* 1 cos a -b a~ l * Ci sin a -{- c* cos a, 
a - " 1 em a ~ C4 cos a — c* sin a. 



Fig. 2 



761 


SEC, 3] THE LAPLACE TRANSFORM 

Hence Ci ** o'* 1 sin a, c& ** a^coe a — 1), and our solution is 
y ~ 0, t < 0, 

y «* n~ l (l — oos t), 0 < t < a, 

y » a~ l sin a sin t — cT^fl — oos a) cos t, a <t 

Sinoe a"* 1 ^ ~ cos t) < o“ x (l ~ cos a) 0 as o 0, and since a” 1 sin a — ► 1 as 
a —» 0, we see that letting a — » 0 gives the solution which was obtained previously by 
the method of Laplace transforms. 

Although h(x) is often called the “Dirac delta function,” it is not a 
function. Indeed, we have already observed that L (/) — ► 0 as p — ► qo for 
every function /, and 8 does not have this property, because L($) « 1. 
It is possible to generalize the concept of function and to generalize, cor- 
respondingly, the definition of L. The process leads to a branch of mathe- 
matics known as the theory of distributions. 1 In this theory manipulations 
with 8(x) of the type carried out in the foregoing discussion are fully 
justified. 

Although a brief and correct definition of the unit impulse 8(x) is not 
easily given, it is easy to define what is meant by the response of a system 
to the unit impulse. Namely, find the response to the function 5 a (x), as 
in the foregoing example, and then let a — ► 0. The Laplace transform 
gives the result of such a calculation directly, without introduction of 
8 a (x). 

PROBLEMS 

1. The voltage V of a certain circuit satisfies 

V" + 4V' + 3F - E(l) 

where E is the applied voltage. Find the response of the system to a unit impulse at 
t m 0 if V »» 0 for ( < 0. 

2. (a) Solve the equations 

y' - «(*), y" - «(*), V m - B(x) 

assuming that y » 0 for x < 0 and that ij and a& many derivatives as possible are con- 
tinuous. (6) Show that y, y', and y " have a jump of value 1 at x « 0 in the three cases, 
respectively. 

3. A certain function U{x) satisfies 

aHJ" -p 2 U - -**«(*), * >0 

where a and 0 are positive constants. It is known further that 17(— x) U(x) t U is 

continuous, and U 0 as x — ♦ <*>. Obtain the solution 

U « (2 *0)“V W « )W . 

Hint : In forming L(U"), take U(0) * c } U'{ 0) = 0 where c is a suitably chosen constant 

l L. Schwartz, “La th^orie de distribution,” Hermann & Cie, Paris, 1950, See also 
B. Friedman, “Principles and Techniques of Applied Mathematics,” chap. 3, John 
Wiley & Sons, Inc., New York, 1956. 



APPENDIX 


[APP. 8 


782 


4. The singular solution u(z,() for heat conduction satisfies 
tt< - oVm, u(*,0) - >$«(*)• 

(a) If U(x,p ) — L(u), where the transform is with respect to t, show that <**17** — pV 
m — (6) Using the result of Prob. 3 followed by Table 2, deduce that 


u(*,t) 


-—x®/ (4a*<) 

2«M)* 


/7raL* The role taken by £ in the table is taken by t in this problem. 


4. Additional Properties of the Transform. The usefulness of the La- 
place transform is greatly increased by the properties tabulated in Table 1. 
Entries la, lb, and 4a were derived in the foregoing discussion, and the 
others will be derived now. To deduce the relation 2 a we have 


[fix - c)e~ px dx = [j(t)e- p(t+c) dt = e~ pc [j{t)e~ pt dt 

upon setting t = x — c. The limits (— o) can be changed to (0,°o) if 
f(t) « 0 on the interval (~c,0), and 2 a follows. In particular, 2a holds 
if c > 0 and f(t) = 0 for t < 0. The relation 2b is simply the identity 

f *-(->*/(*) dx = (V p V*/(z) dx. 

Jo J o' 

This is valid without restriction on c, provided p is large enough. 

For 3a we let i « cx to obtain 

[f(cx)e~ px dx = [f(l)e~ (p,c)t d Q - ^ F Q 

as desired, provided c > 0. Writing 1/c instead of c in 3a gives 3b, again 
for c > 0. 

The result 4b follows by differentiating (1-1). For 5a we apply 4a to 
the function 

fi(x) - ( 7(0 dO 

Jo 

noting that f x (0) = 0 and that /!(«) = /(x) at points of continuity. The 
result 5b follows from integration of (1-1). 

The convolution theorem t item 6 in Table 1, can be established by the 
following device: Since the Laplace transform involves f(x) only on the 
range (0,<»), we can agree to take fix) = 0 for all negative x , With a 
similar convention for g(x), the respective Laplace transforms may be 
written in the form 1 

, 1 Transforms of the type (4-1) are called bilateral , in contrast to the unilateral trans- 
form (1-1). An account of the bilateral Laplace transform may be found in B. Van 
der Pol and H. Bremmer, “Operational Calculus/’ Cambridge University Press, London, 
1950. 



me. 4} 


THE LAPLACE TRANSFORM 


763 


Uf) - f e-* x f(x) dx, L(g) - f* dx, (4-1) 

*—90 * —00 

and the function A(x) of Table 1, entry 6, is equal to 

h(x) = r f(t)g(x ~ k) dt (4-2) 

* — 00 

Indeed, the lower limit — » in (4-2) may be replaced by zero because /({) 
* 0 when £ is negative, and the upper limit may be replaced by x because 
g(x — £) = 0 when x — £ is negative. Given (4-1) and (4-2), the convolu- 
tion theorem L (h) = L(J)L(g) can be proved by a discussion which is 
practically identical with a discussion given previously, and hence we do 
not repeat the argument here. 1 

Example: Periodic Functions. Let Pq(x) be the function illustrated in Fig. 3, bo that 
Po(x) « 1 for 0 < x < a, P 0 (x) = 0 elsewhere. 

Direct computation gives the transform 

LlPo(x)] - /* e~**dx « p“Hl - e~* p ). 

J o 

If the function is translated c units to the 
right, as shown in Fig. 4, the result is 

h[P 0 (x - c) 1 * p~\ 1 - e-* p )e~ pc (4-3) 

by Table 1, entry 2 a. Upon choosing c » 0, 
c m b, c » 26, c » 36, ... and adding, we get 
a square wave * y **= P(x). According to (4-3) 
the Laplace transform is 

1 _ f ~-ap 

UPix)} - P~'{ 1 - «-° p )(l + e-** + e~ lpi + e- Jp6 + • • •) - 
when we recall the formula 1/(1 — r) for sum of a geometric series (Chap. 2, Sec. 1). 




1 See Chap. 6, Sec. 18. In the present case the integrals are absolutely convergent, 
the change in order of integration is justified, and the prooess actually gives a valid 
proof. 

* See Fig. 5. It is left for the reader to sketch the graph when b «■ a and when b < a. 


764 appendix {app. b 

The procedure just described can be applied to any periodic function P(x) and yields 
the formula 

L[P(*)] « (1 ~^~ l L[P 0 (*)] 

where 6 is the period and where Po(x) * P(s) on 0 < % < b but Pq{x) » 0 elsewhere. 



For example, the reader can verify that the transform of the sawtooth wave shown in 
Fig. 6 is 

(1 - e-**>)~ l ab- l p-*l 1 - e-**(l + bp)). (4-4) 



PROBLEMS 

1, Find a function f(x) whose transform is 

2p — 5 

3 p* + 12p + 8 

Hint: By completing the square the expression can be written in the form 

2p - 5 2(p 4- 2) - 9 2 p + 2 3 

3(p + 2)* - 4 " 3(p + 2) 2 - 4 “ 3 (p *f 2)* - % (p + 2) 2 - % 

Use Table 2, entries 2a and 2b, with a «* 2i/y/& (see also Table 2, entry 7). Then 
use Table 1, entry 2b, with c » —2. 

2 . As in Prob. 1 find a function whose transform is 

8. Derive Table 2, entry lb, from Table 2, entry 4a. 

4. Derive Table 2, 5a, from Table 2, 56, and Table 1, 46. 

5, Derive Table 2, 46, from Table 2, 2a, and Table l, 56. 


765 


BBC. 5] THE LAPLACE TBA.NSFOBM 

6. If f(x) m **~t ami g(x) «■ x*~ l , show that 

/ * g - x a+i ~ l / V*‘( 1 - <)‘~ l di. 


Hint: Let £ «* to in the definition of f*g (Table 1, entry 6). 

7. From Prob. 6 deduce the Euler formula for the beta function 



f— *(1 - O*- 1 di 


(a ~ \)Kb - 1)1 
(a + 6 - 1)1 


Hint: By the convolution theorem applied to the result of Prob. 6, 


6. Steady-state Solutions. The Laplace transform will now be used to 
solve the general linear equation with constant coefficients, 

y {n) + On-iV (n ~ l) H 1" «i y' + aoy = f{x) } (5-1) 


subject to the initial conditions 

2/(0) * 0, y'( 0) « 0, . . 0) - 0. (5-2) 

The solution satisfying (5-2) is called the steady-state solution of (5-1), 
because in many physical problems the effect of the initial conditions decays 
exponentially as x increases. 

By repeated use of Table 1, entry 4 a, 

= p k L(y), (5-3) 

for k = 0, 1 , 2, . . . , n, provided (5-2) holds. Hence the transform of (5-1) 
yields G(p)L(y) = L (/), or 

Uv) = — L (/), (5-4) 

G(p) 

where G(p) = p n + a n _ 1 p n_1 -| b ajp + a 0 . 

Determination of y from (5-4) is especially easy when G(p) has only simple 
roots pic 0. Indeed, expanding 1 /G(p) in partial fractions leads to 

L(l/) = L(/)S — — - — (5-5) 

V -Vk 

where the A *- s are constant. 1 Since Table 2, entry la, gives 
2 — = 2A*L(e p **) = L(2A*e p **), 

V ~Vk 

Eq. (5-5) may be written 

L (y) - LCf)L(2A*e”**). 


1 If we multiply through by p — p* and let p — * p*, it is found that 1/A* - O’iPk)- 



766 APPENDIX (app. b 

Comparing with Table 1, entry 6, gives Heaviside's expansion theorem 

V J df. (6-6) 

The essence of the method is that (5-4) leads to 

Uv) - Uf)UQ) 

provided g is a function such that 1 /G(p) = By the convolution theorem, 

v - / * o - - i) dt 

This formula is valid even when G(p) has multiple roots, though the determination of 
g may then be more difficult. 

The function g can be thought to be the steady-state solution of 

9 {n) + a»_i0 (n “ l) 4 j- aig f -f aog « 6(x) (5-7) 

because the transform of (5-7) yields G(p)L{g) * 1. However, since we have not de- 
veloped the theory of distributions, it is better to avoid the use of S(x). This question 
will be discussed next. 


Let h(x) be the steady-state solution of 

A (n) + a n _ x h {n ~ l) + • • • + a, A' + aoA = J(x) 
where I (x) denotes the Heaviside unit function: 

I(x) = 0 forx < 0, I(x ) = 1 forx > 0. 

The Laplace transform of (5-8) yields G(p)L(/i) = 1/p, so that 


L(h) 


1 

pG(p) 


Writing (6-4) in the form 


Uv) - pW) 


we obtain 


pG(p ) 


L(p) = (L(D +/(0)]L(A) 


= pL(/)L(A) 
L(f')L(h) +f(0)Uh) 


(5-8) 


by Table 1, entry 4o. The convolution theorem now yields 

V - />«)*(* - i) di + /(0)A(x). (5-9) 

Thus, the steady-state solution of (5-1) can he obtained from the steady-state 
solution of (5-8) 1 by means of the formula (5-9). This important fact is 
known as the Superposition principle . 



767 


SBC. 6] THE LAPLACE TRANSFORM 

As in the derivation of (2-4) one can show that 

L (y<*>) - p k h(y) - p*~VO) - p*~V(0) y (k ~ l) (0). 

By means of this formula the Laplace transform can be used to solve (5-1) subject to 
the general initial conditions 

1/(0) - yo, V'(0) -i/i, . . y (n ~ l) (0) - 

It should be emphasized, however, that the superposition principle applies to steady- 
state solutions only. 


PROBLEMS 


1. Find the steady-state solution of 

y" 4- 3y' + 2y -/(*) 

by use of Heaviside’s expansion theorem. 

2. Evaluate the result of Prob. 1 explicitly when 

(«)/(*) - I(x), ( b)f(x ) - e a *. (c)f(x) - z. 

3. By means of the superposition principle obtain the solutions (b) and (c) in Prob. 2 
from the solution (o). 

6. Integral Equations. An equation of the type 

ff(x) = \f(x) +f Q f{Z)k{ x - £) d£ co-l) 


where X is constant is called an integral equation. It is supposed that g 
and k are known and that / is to be found. Because of its close relation 
to the convolution theorem, this equation lends itself to analysis by means 
of the Laplace transform. Indeed, taking the transform of (6-1) yields 


L (jg) - XL (/) + L(/)L(*) 

when we use the convolution theo- 
rem. Hence 

L (g) 


Uf) 


X 4" L(/c) 


and from this, / can often be found. 

The process will now be illustrated 
by an example. 

Starting from rest, a particle slides down 
a frictionless curve under gravity (see Fig. 

7). It is required to determine the shape of 

the curve so that the time of descent will be independent of the starting point. 
A curve of this sort is called a tauiochrone . As we shall see presently, the only 
tautochrones are cycloids. 1 



1 For another interesting property of the cycloid see Chap. 3, Bee. 14, Prob. 3. 



708 APPENDIX [APP. B 

If tike particle starts at a height y, its velocity » when the height is y\ «• e can be 
found by equating potential and kinetic energies. The result is 

yimfi — mg{y — ij) or v ** (2g) *(y — rj)^ (6-2) 

where g is the acceleration of gravity. Denoting the arc along the curve by s, we see 
that the time for descent is 

/ t-O -/;«•>* 

where f(ij) stands for ds/dy at y * Since the timeis constant and since t> is given 
by (6-2), the problem reduces to 

J^fWiv - v)~ H dv - co 

where cq is constant. This is an integral equation for /. 

Taking the Laplace transform gives « L(co), or 

- cop-' 1 . 

This gives t(f) ■» Cip~ H , where ci is constant, and hence f(y) « cy~ H , where c is con- 
stant. Thus we are led to the differential equation 

»-5-[' + ©T-^ 

If we set y ** c 2 sin 2 yfa> t a short calculation yields 

x m + sin 0), y « - cos <fi) 

which are the parametric equations of a cycloid. 


REVIEW PROBLEMS 

1. The current I in an RL circuit satisfies 

L d l + Rl-V 

where V « V(t) is the applied voltage. At time N 0 a switch is closed, so that V 
suddenly assumes the value Vo -f A sin U. (Here L, R, Fo, A , and w are constants.) 
By use of the Laplace transform find I for t > 0. 

2. Find the response of the circuit in Prob. 1 to a unit impulse at time t « 0, assuming 
that V * 0 for t < 0. 

3. Find the steady-state solution in Prob. 1 when V is an arbitrary function by (a) 
the Heaviside expansion theorem, (b) the superposition principle. 

4. If L (y) « F(p) use Table 1, entry 4b, to obtain 

Uxy) - -F', Uxy') - ~(pF)\ L(*y") - ~(p 2 F)' + y(0). 

6. A function y satisfies xy" -f V' + xy • 0 and has a Laplace transform L(y) ** F(p). 
By use of Prob. 4 show that 

F'(l + p 2 ) - -pF, 
and thus deduce that y — cJv(x) where c is constant. 



sue. 6] 


THE LAPLACE TRANSFORM 


0. An insulated rod extending along the positive x axis is initially at temperature 0, 
and the end x — 0 has the temperature /(f) at time L The temperature u(z,t) satisfies 

u, - a*u„, u(x,0) - 0, u(0,<) - /((). 

(o) If U(x,p) is the transform of u with respect to t, show that 

U - L </) e -VP*/«. 

[Assume that U — ► 0 as x — * and note that U(0,p ) ** L{/).] 

(6) By writing U « L(/)L(^), where p is found from Table 2, deduce that 


u(x,t) 


2a\/ 


rq-f<j)dr. 


J r* e ~x*/i4a*{t- 
o 0 — t)* 


7. Use the Laplace transform to solve some of the text examples and problems in 
Chap. 1, Secs, 21 to 26. 


Table 1. Phopekties of L[/(x)] ** F(p) 


i 

; 

a 

b 

1 

L(/ + (?) - «/) + LO?) 

Ucf) - cb(f) 

2 

L[/(x - «)] - e~ pc F(p) 

F(p - c) - He'*f(x)} 

3 

Ll/(cx)J = (?) 


4 

L[/'(*)] - pf’(p) -/(0) 

F'{p) - L\ — x/(x)J 

5 

1 [ r /(< > ^ 


6 

L(/)L(p) - L(/i) where A(x) - f/(£)s>(x - £) d£ 

^0 












APPENDIX C 


COMPARISON OF THE RIEMANN AND 
LEBESGUE INTEGRALS 


1« The Riemann Integral. Let a function fix) be given on the interval 
a < x < b (Fig. 1). To define the Riemann integral 

fj(x)dx (1-1) 

we divide the interval [a,b] into smaller intervals by points x*, 
a = x 0 < Xi < x 2 ■ • * < x n = b. 



It will be desirable to consider a sequence of subdivisions which are made 
finer and finer by choosing more and more points x*. The precise require- 


ment is 


and 


max 

k 


X k ~ Xfc-1 1 


0 . 


To describe this situation we say, in brief, that the subdivision becomes 
arbitrarily fine. 

Let be an arbitrary point on the interval With yk » /(£*) 

as shown in Fig. 1, the sum 

s = Vl(*l - so) + Vnfa - *l) H 1- Vn(x n ~ X»_l) (1-2) 

771 



772 


APPENDIX 


[APP, 0 

represents a certain area that presumably approximates the area under the 
curve y ** /(#), The geometric interpretation suggests that s has a unique 
limit independent of the manner of subdivision, provided the sub- 
division becomes arbitrarily fine. When s 
actually does have this behavior, f(x) is 
said to be Riemann integrable , and the 
limit s 0 is called the Riemann integral . 

The Riemann integral does not exist if /(x) 
oscillates too violently. For example, let o 0, 
b » 1 and define 

f(x) » 2 for x rational, 1 

(1-3) 

fix) ** 3 for x irrational. 

It is easily shown that every interval (no matter 
how small) contains both rational and irrational 
numbers, so that the graph of /(x) has the appear- 
ance suggested by Fig. 2. If we choose & rational, 
then /(£*) = 2 and 

8 « 2(xi - x 0 ) + 2 ( x 2 -xi) ■+•••• 

-f 2(x„ - x n _i) « 2(x» ~ x 0 ) = 2 

no matter how fine the s\ibdi vision may be. 
On the other hand, if the bus are all irrational, 
then This shows that the limit of 8 depends 

on the maimer of subdivision and, hence, that the 
Riemann integral does not exist. As we shall see 
presently, the Lebesgue integral for this function 
does exist and can be evaluated explicitly. 

2. Measure. The decisive idea in the Lebesgue integral is the notion 
of measure , which will now be described. The measure of an open 5 interval 
a < x < b is simply the length b — a. If a set consists of a finite collection 
of such intervals (Fig. 3), the measure is the sum of the lengths. The 



Fig. 3 

same definition applies when there are infinitely many intervals. The 
sum of the lengths is now an infinite series, but since the terms are positive, 
the mm does not depend on the order of the terms (Chap. 2, Sec. 6, Theorem 
III). Thus, the measure is well defined in this case also. 

1 A rational number is a fraction p/q where p and q are integers. Thus and ~~ l % 
are rational, but y/2 is not. 

* An interval is open if the end points do not belong to the interval and dosed if they 
do. Thus a < x <. b is a closed interval. 




RIEMANN AND LBBESGTTE INTEGRALS 


TO 


SBC. 2] 

The notion of measure can be extended to still more general sets E 
as follows. Let / be a collection of open intervals which contains 1 E, 
and let m{I) denote the measure of I. We approximate E better and better 
by these sets I, so that m(I) becomes smaller and smaller. The smallest 
value for m(I) which is given by this process is called the outer measure 
of E and is denoted by 

Strictly speaking, the “smallest value” need not be attained, and the precise definition 
of outer measure is as follows: The outer measure is the largest number c such that 
m(7) > c for all sets I of the above-described type. The number c is called the greatest 
lower bound of the numbers rn(I) ; its existence can be established by the fundamental 
principle quoted in Chap. 2, Sec. 1 . 

A collection of open intervals, such as 7 in the foregoing discussion, is 
called an open set. As we have seen, outer measure is defined by consider- 
ing the open sets containing E. The points of [a } b] not belonging to a given 
open set form a closed set. By considering closed sets contained in E one 
can define the inner measure m t (E). If m,{£) = m 0 (#), the set E is said 
to be measurable and the common value is called the measure of E. 

To illustrate the calculation of a measure, let the set E consist of the rational points 
x on 0 < x < 1, that is, the points whose coordinate x is a rational number. By taking 
first the rational numbers p/q with denominator q — 1 , then those with q m 2 , and 
so on, we see that the rational numbers can be arranged in a sequence 

n, r 2 , r 3 , . . ., r n , .... ( 2 - 1 ) 

Given « > 0 , construct an open interval of length e/2 centered at r h an interval of 
length «/ 2 2 centered at r 2 , and so on. Th** nth interval is of length e/ 2 " and is centered 
at r n . If / denotes the set consisting of all these open intervals, then 

m(/) <-+£ 5-1 ~ H ( 2 - 2 ) 

[We have inequality rather than equality in ( 2 - 2 ) because some of the intervals may 
overlap.] 

The foregoing construction shows that the outer measure of E is < e. Since « is 
arbitrary, the outer measure must be zero. Because m x {E) S mo(E), it follows that the 
inner measure is also zero and, hence, m(E) » 0. 

As a second illustration we shall find m(E f ), where E' is the set of all irrational numbers 
on [0,1]. One of the most important properties of measure is that it is additive ; if E 
and E f are two measurable sets with no point in common, then 

m(E + E') - m(E) + m(E'). 

(We use E + E' as an abbreviation for the set of all the points belonging either to B 
or to E\) In the present case E is the set of rational points on [0,1], and E f the set of 
irrational points on [0,1]. Evidently E + E f is the set of all points on [0,1], so that 
m(E -f E*) * 1 . The above equation then gives 

m(E') m l — m(E) =*1—0*1. 

1 That is, every point of E is interior to one of the intervals belonging to the set 7. 



APPENDIX 


774 


[app, c 


S» The Lebesgue Integral. A function y * fix) is said to be measurable 
if the set of points x at which fix) < c is measurable for any and all choices 
of the constant c. It can be shown that the set e* at which yk~i < fix) < yu 
is then measurable for all choices of yk~ 1 and yk< To define the Lebesgue 
integral of fix), let the y axis be subdivided by points yk as shown in Fig. 4, 
and form the sum 


a * yimiei) + 2/ 2 m(e 2 ) H b 2/«m(e n ). 



When fix) is measurable and bounded, the sum a has a unique limit <tq, 
independent of the manner of subdivision, provided the subdivision be- 
comes arbitrarily fine. This limit a 0 is called the Lebesgue integral of fix) 
and is written in the form (1-1). 

The most obvious difference between Riemann's definition and Le- 
besgue’s is that in the former the x axis and in the latter the y axis is sub- 
divided. This distinction, however, is superficial. The important fact is 
that Riemann’s definition is based on the notion length of an interval 
whereas Lebesgue’s is based on the more general notion, measure of a set . 
The intervals x* — in Riemann’s definition play the same role as the 
sets e* in Lebesgue ’s. 

Riemann's definition breaks down if fix) does not remain close to yk 
throughout most of the intervals [&*_!, x*]. Lebesgue J s definition cannot 
break down in this way, because fix) is automatically close to tjk through- 
out the set e*. That is why (in contrast to the former definition) the 
latter carries with it an assertion that the integral actually exists. 



RIEMANN AND LEBESGUE INTEGRALS 


775 


SEC. 3] 

To illustrate the calculation of a Lebesgue integral we shall integrate the function 
(1-3) illustrated in Fig. 2. If the intervals (y p -i,y p ) and (yq~i,y Q ) contain 2 and 3, 
respectively (Fig. 5), then the sets e p and e q are the only ones that are not empty. Thus 
m(cjk) ~ Q for k 9 * p or q, and the sum reduces to 

<r « y p m(e p ) *f y g ni{e g ). 

Since ep is the set of rational points and e g the set of irrational points, these sets have 
the measures 0 and 1, respectively. Hence <r * y q . As the subdivision becomes 
arbitrarily fine, y Q — ► 3 and the Lebesgue integral is found to be 


f/(x) dx - 3. (3-1) 

Jo 

It can be shown that if the Riemann 
integral exists, then the Lebesgue integral 
exists also and the two have the same 
value. On the other hand, the latter may 
exist when the former does not, as we 
have just seen. Because of its greater 
generality the Lebesgue integral has 
many desirable properties, of which we 
mention the following: 

Lebesgue Theorem on Bounded Con- 
vergence. Suppose j/ n (x) | < M where 
M is constant , suppose f n (x) are Lebesgue 
integrable , and suppose lim f n (x) — f(x) o?i 
an interval [a f 6]. Then fix) is Lebesgue 
integrable , and 

rb rb 

lim / f n (x) dx= f(x) dx. 

Ja Ja 

To see why the theorem fails for Riemann 
integrals, let / n (x) « 2 at the first n rational 
points rk in the sequence (2-1) and f n (x ) » 3 
elsewhere. Then |/ n (x) | < 3, and as a Riemann 
or Lebesgue integral, 

fuix) dx - 3. (3-2) 

Jo 



Fig. 5 


Evidently, lim/ n (x) ■» f(x), where /(x) is the function (1-3). Taking the limit of the 
expression (3-2) as n — ► », we get 

lim Cfnix) dx - 3 - Cj{x) dx (3-3) 

Jo Jo 


provided the latter integral is the Lebesgue integral (3-1). Equation (3-3) does not 
hold for Riemann integration because, as we have seen, f(x) is not Riemann integrable. 



APPENDIX D 


TABLE OF *(*) * ~= f* it ' 


x 0.00 0.01 0,02 0.03 0,04 0.05 0.06 0.07 0.08 0.09 



* This table is reproduced by permission from the “Biometrics Tables for Statisti- 
cians,” vol. 1, 1954, edited by E. S. Pearson and H. O. Hartley and published by the 
Cambridge University Press for the Biometrics Trustees. 

776 




















ANSWERS 


Section 1, Pages 10-11 

1. Ordinary, fourth order. 

4. Ordinary, first order. 

7. Ordinary, second order. 

14. y *» x 2 ;y « x 2 ; y « x 2 

x 8 x 8 

15. y _ -g- -f x;y - — + t 


CHAPTER 1 

2. Partial, fourth order, 

5, Ordinary, second order. 
8. Ordinary, third order, 
-f 1 ; y « x 2 — 2. 

JlX + C2. 


3. Ordinary, first order. 
6. Partial, second order. 


Section 3, Pages 13-14 


1 . 3 ^. 


3. p *» ~/c; a * 

6. v * 30e“*‘; » 

7. 2 in. 


- k). 

- J ( 1 -•-*>. 




4. (log 2)/2. 


6. 2 32 hr; qo. 


Section 4, Page 16 

2. The rate g would be thought of as/(0<7 instead, where /(<) « 0 for 0 < t < and 
/(f) « 1 for f > fo* Equation (4-5) would be written dx/dt » wr *-f- /(f)<? — rx/p. 

3. dx/dt * wr + fcAoe"* 4 - rx/j/. 6. (A - x)/(# - x) - [A/B)^ a ^ b \ 

6. Let x represent amount of substance dissolved after time f, A the amount of sub- 
stance present when x ** 0, t «* 0, and c the proportionality constant. If v is the 
volume of solvent and S the saturate concentration, then dx/dt ** c(A — x) X 
(S — x/v) if the dissolving substance docs not change the volume v of the solvent. 


Section 6, Page 18 

1. sin”- 1 x — sin™ 1 y *» c. 

3. 2 cos jy — sin x cos x -f- x «** c. 

6. (tan™ 1 y — 2VT -f x) I * 0. 

1(0,1) 


Section 7, Page 20 

‘■[K ,+ ?)]’ 


3. sin - *f log x 
x 


c 

x 


2. ( y — J)/(y + 1) 588 cc T ®. 

4. (sec x — tan y)| » 0. 

*( 0 . 1 ) 


6. ( y + l)/(x + 1) - 2. 


2. sin*” 1 - — log x ** c. 
x 

4. i* - 2* V - y 4 - -2. 


777 



778 


ANSWEBS 


*• logy 4 - 


6. y - x 4* y log x - 0. 


7. 2 tan”* 1 (e*0 4* log tanh ; 

* 

9. * - cs*^. 

11. log x 4 «~ v/ * « c. 

Section 8, Page 22 

1. a* 4 * 4 y ■ c. 

4. xy « c. 

7. Not exact. 

10. Not exact. 

Section 9, Page 28 

7. y/z 4 x — c. 


8. y - ee~ 2Vxlv . 

10. — — ~ — log y « c. 
x V 

12. y(2 - log y) = ^ tan 2 x 4 c. 


2. Not exact. 

5. sin (y/x) *» c. 
8. x 2 4 sin xy » 


8 . x — eye 11 


10. y — 2 tan - " 1 (x/y) « c. 11. (x/y)e v 
Section 10, Page 26 

1. 1 + Vac* + 1 - cxe-*' 1 ' 3 *'. 
a. y - e~**(x - 1). 

5. y « cos 2 x 4 2(sin x — 1). 


8. y ■» sin x 4 ce*. 

10. x - 1 4 ce"^ 2 . 

12. y - 

12. x sin"" 1 x 4 V7 - x 2 
Section 11, Page 27 

1. y * cie 8x ; y * cyT*. 

p 4 — 2 

S. y ~ — ; x - 2p 
p 


3. x 8 y — xy 8 * c. 
6. x 2 4 y 2 ** c. 

9. x 2 y 4 xy 2 4 x 


9. yx* - ce v . 

12. y 4 x 2 /y * c. 


o 

• V “ 3(x* + 1)' 

4. y - 1 - 2e“*'* 

2 , 

6. y » 2 sin x ~ x cos x 4 - cos x 4 * 

9. x - ce w,) ' / * = *. 

11. x(l + 4y 2 )* - c. 


- - ]) - o. 


2. y - x + cj; y - x + 


4. V - log (p* + 2p); x - - 1 tan 1 ^ + c. 

(^X 4 1 f~ . 

6. y «* ; y *» 2 Vx. 6. sin“ A y d= x « c. 

c 

7. y * cs x ; y « c — x 2 . 8. y - e* 4 c; y ■* c — x*/2. 

Section 12, Pages 28-29 

1* y 4 ** (48x“~ 2 — 96x“* 4 — 4) coe x 4 (16x _1 — 96x~ 3 ) sin x 4 cx” 4 . 
2.f 2 -*4«4 c#\ 8. y“ B - M* 3 4 cx*. 

4. x «■ y log cx. 6. y~ l ** 1 4 log x 4 cx. 

6. y~ 3 — 1 + x* + ce* 1 . “ + ” - 


8 . 3 

ax 


7 » — — ;i-u~3,y-p4i 

aw 2w 4 tf 


■ ; u * 3x 4 y 4 7. 


. u — 0 

> srn — ■ — ; x 
w 4 » 


u — y ** v — 

2 2 



ANSWERS 


779 


10 . 


du 


— I m cos u; u W x 4- y. 
1 y a 4- cy. 


dx 
12. 2x 

14. x see y ®» log | sec y 4* tan y | 4- c. 


11. x~ 2 + «** 


16, tan"* 1 y — tan“ 
19. - c - x. 

23. y - e 8x + ce 2x . 


x ** c. 17. 4* y 4 


13. y * log (ary - c). 
16. y tan" 1 
** c. 


20 . 

24. 


- y 4- H 4- cc 2v * 
2x 2 c w + cx 2 . 


x » c. 

18. xsin2y 


21. 4x « 2y - 1 4* cc~ 2v . 


Section 14, Page 33 

1. y » cx, 2. at 2 — y 2 »» c. 

8. x 2 4- ny 2 « c. 6. 0 * c. 

9. Self-orthogonal family. 

10 . x 2 — 2ax 4- y 2 ** 0; 2xyy' 4* x 2 — y 2 ** 0; a family of orthogonal curves is 
x 2 - 2 ay 4* y 2 * 0. 


Section 16, Page 34 

1. (a) y * cx; (6) x 2 4- y 7 «* r 2 . 

2. (a) o ± x - VT- y* - 1«K 1(1 + VT - j 7-)/y}; ( 6 ) y - re'. 

3. (a) y ** c x ; (b) y = cosh x. 

4. - c(i + Vx^+'^r 6. t - |x 0 |6/(fc 2 - a 2 ); hi -To 1 - «. 

Section 19, Page 46 

1. t » 100 /y sec. 

6. tan 6 ** tan 0 O 4* 2eE/{(*>mvc). 


Section 20, Pages 49-60 

2. v » V2y/i, s « sin 0. 

4. t> - Vj(l - - 1 


6. y 


2vq cos 2 a 


x 2 4- x tan a. 


e 2k»g/v> * 


Section 21, Page 64 

2. y » — e x 4- e 2x ; y » 0. 4. y «* — 2x 2 4- 4x. 


6. y °» e 2x 4“ 2xe 2x ; y « 0. 


Section 22, Page 66 

1. y ® cie"" 9 ® 4* C2€ 6x . 
4. y * cie 2 * 4" C 2 e~~ 2x . 
6. y «* cic 2x 4* cjxe 2 *. 


2. y ** Cje 3i 4- C 2 C 2x . 3. y » cie* -J- cjxc*. 

6. y * rj cos 2x 4~ C 2 sin 2x. 

7. y ® cic 2r cv)8 x 4* C 2 « 2x sin x. 


Section 23, Page 68 

1. y *® ci«“ x 4- C 2 e x/2 . 2. y 

4. y ■* cie ?x 4- c 2 xe Sx 6. y 

Section 24, Page 63 

1. y m d« 3x 4- C& 2x 4- V 4 *- 
3. y °« cje“" 8x 4“ c#"" 2 * 4- MV** 

6. y « cie* 4* C 20 "~* — 5x 4- 2. 

7. y •» (ci 4“ C 2 x)e x 4- xV/6. 

9. y • ci 4- 4* x/3. 


rje x 4- ^ 2 C x . 3. y ** cie 2x 4- C 2 e*“*. 

cie“ x 4- c 2 X€” x . 

2. y « (ci 4- c 2 x)«- x 4* x ~ 2. 

4. y *=* (ci 4- csx)^ 4- x -h 2. 

6. y - cie x 4- C 2 <r x 4* c 2x (x/3 ~ K). 
8. y «* cje^ 4* cjxe 8 * 4- xV x /2. 



780 


10. y 

11. y 
13. y 

H. v 

16. y 

18. y 
20. y 
22 . y 


ANSWERS 

9x 2 - I8x -f 7 

«* ci sm 3x 4 ca cos 3x H 

81 

« cie x -f cje*"* -f xe x /2. 12. y * 

» Ci« 8x 4 esc 2 * 4 c 2x f x 5 — 3x 2 — 6x 



/X 3 x 2 \ 


■» C1«* + Cj £«* + 

( 6 ~ 2/ 

16. y 

* 3 

13x 2 24x 


- Cl + — - - 

25 + 125* 

17. y 

5 



- 1 - e~ x + x 2 - 

X. 

19. 1/ 

- (3i - 4)/9. 


21. y 


* 2e“ x - 5<r s 79 4 x/3 - %. 


d sin x 4 ca coa x 4 x 8 ~ 5x. 


)■ 


cic 2x 4* cac 8x 4 


2x 2 4 6x 4 3 


-4e~ x 4 2*~* c 4 2e*. 


0. 

0. 


Section 26, Page 66 

1. y « cic x 4 C 2 € 2x — (3 sin 2± 4 cos 2x) /20. 

2. y «« ci sin 2x 4 C 2 cos 2x — (cos 3x)/5. 

8. y ** cjc 1 4 C 2 «"~ x/2 4 2 sin x. 

4. y « ci«" 2x 4 c*r Sx 4 3xe“ 2x 4 c s 730. 

8. y « 6“ x (ci sin 2x 4 c 2 cos 2x) 4 e x (>£ 0 sin 2x — 34 0 cos 2x). 

6. y « cie 8x 4 — 34« 3x cos 3x 

7. y * cie~ 6x 4 c 2 c“ 6x 4 xc Bx /10 - x 2 /25 4 4x/25 - % 2 &. 

8. y » ci sin x 4 c 2 cos x 4 ?& cos 3x — sin 2x. 

9. y *■ — c®/4 4 e~’*/4 — 34 sin x. 

10 . y * 0. 

11. V * -34 4 %oe 2x 4 c” x (Ko cos x 34 sin x). 

12 . y - — cos x 4 (x/2) sin x 4 1 . 


Section 26, Page 70 


1. V 

2. y 
8 . 2 / 
4. y 

6. y - 
6 * 2 / ! 

7. y- 

8* If ■ 

10. y - 


1 ci 4 C2 e~ 3x/2 4 cse 8x . 

1 ci sin x 4 c 2 cos x 4 c~" x (c3 cos 2x 4 C4 sin 2x). 
C\e~ x 4 c&xT* 4 C3X 2 e~ x . 
cie~ 2x 4 czc x sin (\/3 x) 4 cge 1 cos (\/3 x). 

(ci 4 C2x)e x 4 c 3 . 

(cj 4 c 2 x 4 cgx 2 )er r 4 c 4 . 

1 ci cos kx 4 c 2 sin kx 4 C3 cosh kx 4_f4 sinh kx. 


■ ci 4 c* /2 ^c 2 

t e x/Vi ( - 


Vl5, 1 . Vl5 

cos — - — x 4 C3 sm x 
2 2 > 


4 — 4 - 4 


■( 


ci sm - 


2 


^ 4 e~ x/N/2 


4 

V2 

2 


Q sin x 4 C4 cos 


VS 


4 2 cos x. 

11. y » ci<? x 4 c#? 2 * 4 c 3 xc 2ar - x 2 /4 — x — *34* 

12. y « cie""* 4 esc* 4 4H 4 xe r /2. 

18. y - 4 17c x /12 4 7e~~72 - e~ 3 74 - 2x* 4 2x - 5. 


Section 27, Page 72 

1. Dependent. 

4. Dependent. 

7. Independent. 


2. Independent. 

6, Independent. 


8 . Independent. 
6. Dependent. 



ANSWERS 


781 


Section 28, Pages 78-78 

1. (a) y ■» (a? 3 — v?)/Z — 2x/9 — (6) y « e x /12; (c) j/ « x -f 2; (d) 3/ «• ain 

3. y «* — (a^logaO/O. 4. y *» cjc* 4- c?x -f a: 2 4 1. 

Section 29, Page 77 

8 . v' ~ c/[x 2 ( 1 - x 2 )]. 

Section 80, Page 79 

1 . V “ W' 2 + cxx" 1 ± % log x - 

2. y *** cix 2 4* C2X <5 " , “ v/21)/2 4* cgx^™^ 21 ^ 2 — 

3. 2/ - c^i+vT*)/* + ^(iWTo/2 + *2/3. 

4. j/ « cix 2 4- c 2 x — xl(log x) 2 /2 4- log xj. 

5. t/ » c x x 2 4" c 2 x 8 . 6. y « c x x n 4- C2£~ n ” 1 . 

Section 32, Pages 85-86 

1. y ** 2 cos \/T0 *, ; ?/ * 2 cos VT0 < 4- \/i0 sin VTO 

2tt 

3. ?/ « 10 cos \/245 t. 

5 

4. j/ » 10e“ w (cos >/2^0/ 4 siri \ r 22i)t)\R « 400^245(^68. 

V 220 

5. V - ie0\/2 e~ 6W1 cos ^500* -^) ; V’ « lOOf' 500 ^! + 500\/2 0- 

6 . K - 20^ puih 10,000-s/o t + \/5 cosh 10,000\/5 ()• 

d 2 W 

10. 10 + 10 gy * 0; max y « \/3, total drop 2 4 \/3. 

Section 36, Pages 99-100 

1. y *■ Ci cos t 4- C‘i sin t; x * ri sin * — c 2 cos 

2. /y » Cic* 4- e 2 e~ / ; x » ae* — rtf'" 1 . 

3. y » o f (ri 4- C2O; a* *= e jr (r x 4* c 2 /2 4* c 2 /)- 

4. y = rie* 4- cgc"* 4- 03 cos 1 4* c 4 sin f; x «* fic* 4“ c 2 e"* — C3 cob £ — C4 sin £. 

5. y °* Ci(l 4 \/2 )e y/2t 4 r 20_j~ >/2 )c v/2< ; x = cic^* 4 c 2 e*" v/2 ‘. 

6 . y - cje^ 8 4 ^ _4 4- Hi 

* - r, + c 2 (-- I . 4 17 ) fW-VTi)! + 

9. Cycloid of radius mE/(eH 2 ). 

Section 36, Page 106 

1. yi » c x e~ x 4- c 2 e 8r ; 2/2 * 2(r 2 e 8 * - cie~*). 


CHAPTER 2 

Section 2, Page 118 

3. 1(c), 1(c) for x « 0, =fc?r, d=2*-, . . . ; none in Prob. 2. 

4. 1(6), 1(c) for x 9 * 0, =fcw, db2w, . . . ; 2(a), 2(d). 

5. 1(a), 1(d); 2(6), 2(c), 2 (/).' 7. (a) Yes; (6) no. 



ANSWERS 


Section 3, Pages 121 -122 

1* div, «““* i» r~~» div. 
3 log 2 


2. Con, con, div, con, div. 


3. (a) c < 0; ( b ) con for c < 0. 4. c m < n < e im . 

H < * < %* 1.08190 < s < 1.08267. One term; eight terms. 

Section 4, Page 124 

1. Div, con, con, div. 2. Div, con, div, con, con. 3. Con, div, con, con. 

4. (6) No. For example, «n *■ R» b n «• n, c n «■ 1 — n, d„ « 2 — n. 

log (9 c) 

3. JV > - r— ♦ 
log 10 

Section 5, Page 127 

1. Con, div, con, con for |x| < \/5, con. 2. Con, con for c > 1 only, div. 

3. Con, eon, con, div. 

Section 6, Page 132 

1. Cond con: div, abs con, abs con, abs eon. 

2. Abs con: |x| < 1, all x, |x| < 1, |x| > 1, — ^3 < x < 4, x *»0, \x — 2\ < 1, all x. 
Cond con: x 1, never, never, x = —1, x » never, x «* —3. 

8. 0.95. 

Section 7, Pages 137-138 

3. Unif con for (a) — °o < x < »; (b) |x| < c < 0 1; (c) | (2/n)x — n| > c > 0, 

wher& n is the odd integer nearest to (2/w)x; (d) 1 < c < | x ] < 00. 

4. Unif con for (a) — « < x < (b) \x\ < c < 0.1; (c) 2 x/r ** odd integer or 

{ (2/r)x — nj > c > 0, where n is the odd integer nearest to (2/r)x; 

(d) |x| > c > 1 or |x| < c < 1. 

6. Yes, no, no, yes. 6. No. 

Section 8, Pages 142-143 

1. Con for — 1 < x < 1, — \/2 < x < y/2, all x, — 3 < x < 3, — \/3 < x < v^3. 

5. (b) tan-' x - £ 4. (rf) 2.72, 0.368. 


S. ( b ) tan -1 r •> 2 ~ 2 2n ~ fl . 


6. S(-l) n — — , E(-l)"- 1 — 

4ft -f-1 o (it 4“ 3) (5n. 4- b) 


E jl±i x8 . 


Section 9, Pages 146-147 


, 2"(x - 1)" 


x 2n+1 7r 2n+1 fx — l') 2n + 1 

— — r 2n + l =* ~ ii • 

WZ{ 1} (2n + l)! ( J (2n + 1)! ’ 

Jin (x l) 2n 

<’> ~ : ■«-'>■ «si + “ 1 £ <-»- sttwi - ’-“'ir 

ft)2 + *’ -3+ 2ft- 1) + ft - l) 1 ; 


«£*=!?■ 




i. ^jS-" ; w s(-«- 

ft! ft 4* 1 



ANSWERS 


783 


» (a) 2( — l) n (x - 1)*; (b) - jZ (i + ~^r) (* - 1)»; 

W-gZ(l+2(-l)*-^) (*-«». 

*■ (' - if + 5 ^ ranTijr (* - ;)* 


x 2n .~4n+2 

6.S-.)* - r ,X ( - 1 )- s _ n5j , 


r 2*+l 


2 . V 

^(2n)I ^(2n + l)! 


2(~D n 


n + l‘ 


Section 10, Page 149 


1. (o)S(-l)" 
(c)2(— 1)" 
3. 2(— 1)’ 


^n+l 

(2n *f l)n! 

~4n+8 


;(b)2j2 


(4n + 3) (Sin 4- 1)! 

x p+n 

for p > 0 and all x. 


(2 n -f l)(2n + 1)! ' 

r 2»41 

; (d) 2(— i) B - 


(2n + l)(2n + 1)! 


n!(p + n) 

1 f («? ~ 1)(<? ~ 2) 

' J> i n!(p + n)(-l)» 


(9 ~ «) 


j > 0, p ^ 0, —1, —2, 


Section 11, Pages 162-158 

2. (a) a: + 4- Ks** 4 — ; (6) 1 + x + M* 1 4 — ; 

(c) l 4- 4- M**‘ 4-- • •; (<*) H + H* - Hs** +■■■; 

<e)l-X*-Msa? +•••;(/)! 

8. (a) 0.00133. 4. 4- «** +• • 

6 . 3.004, 0.986, 0.839, 2.036. 6. 0.310, 0.020, -1.025, 0.94. 

7. |«| < 0.24 radian - 14°. 9. H** - 34 2 * 7 - 

Section 12, Page 166 


, 2 ” 


1. Z-fX”, 1 +x 4- 2 1. 


'n! 


3. sin 


ni 

.-1 -8 — ( 2 » — 1 ) 1 
2*4 • • * 2n 2n + 1 


2. y - 1 + a: 2 + Hx*', k « 2. 


r 2n+l 


Section 13, Page 169 

X 2n X 2»*U 

1. (a) ooS( - l) n 4- 0 iS( - l) n ; 

x } ( 2 n) 1 “r®X V ) Qn + iy.' 


(b) 1 -f «o cos x + «i sin x; 

l-4x 6 , 1-4-7* 9 


(c) 4 + h 


2. (o) e*; (6) x - 1. 


61 


9! 

+ • 


+...)+( 

M' 


X 2i' 2-5x 7 , 2-5-8 1 * 

ni i! + ir + ~7r + ~ior 


21 + 61 + ‘ 8! ‘ 111 


8. Ci2(-l) n — 4-e»2* n . 

n! 


)• 


Section 14, Pages 166-166 


6. On ** — 


1 


(n 4- p)n 


On-*. 


7. (b)J—(ci 

> irX 


(ci cos x -f c* sin x). 



ANSWERS 


784 

Section 17, Page m 


2. 


sin (n -f* l)x/2 sin (ruc/2) 


4. 12*. 


sin (x/2) 


X 9* 0, ±2r t db4ir, 


Section 18, Pages 182*18$ 

o «6 /_ iNn-l 0© / _i\» «o 

*•-*■ + £ ^ cog (2n - l)z + E ^r~ sin 2nx + E ( ' 


8' ‘ i 2(2n — 1) 

Section 18, Page 187 

1* e, e, o, e, o, neither, e. 

Section 20, Pages 191*192 

e 4 V 8in ( 2n jl jXg/gjf 

* TO 2n + 1 
4. COS rX. 


4n 


1 


2 T ^ 2 (2n * 1) J 


Section 28, Page 204 

4 4 

2. a x - * ; os ■» 0; a# * — 
r 3ir 


CHAPTER 8 

Section 1, Page 219 

(a) Entire xy plane; ( b ) entire xy plane; (c) y 2 < 5; 
(d) ^ + y 8 ^ 0; (e) i ^ 0; (/) (x - l) 2 + y 1 < 1* 


Section 2, Page 222 

t <«) \ :(&) ** - rdbr.’ l! + 


X* X ' ' ' X 2 + y 2 ' 

(c) y cos xy -f X, x cos xy; ( d ) e* log y, e*/y; 

1 


(e) 2xy -f 


vr 


2* (a) 2xy — **, x 2 + i y y - 2x*; (6) yz + -» xz *f -> xy; 

x y 

t —zx 


(c) 

» 

(•) 


*> sin" 


Vy 2 — x 2> yV y 2 — x 2 * y 

x y t 

Vx 2 + y 2 + « 2 ’ Vx 2 + y 2 + 2 2 V^x^+'y* +'«* * 

— x -y -2 

(x 2 4* y 2 *f e 2 )* 2 ’ (x 2 + y 2 -f 2 2 )^' (x 2 4- y 2 -f « 2 )^* 


Section 8, Pages 227*228 

1. ir/6 ft 8 . 

4. 2,250. 

7. 0.112; 0.054. 

10. l.Oir; r. 


l)«*i ~ ain nx. 
2n 


cos (2n — l)irx. 


2. 11.7 ft. 

5. 10.85 
8. 53.78; 0.93. 


8. 0.139 ft. 

0. 88.64. 

9. 0.003r; 0.3 per cent. 



ANSWERS 


1 


786 


Section 4, Page 280 

1. aa*/a 2 4 yyo/h 2 - 1. 

i ay + V2 a6. 

8. (o) «<* (2t an ^ ooe 

(6) 2r(l — 3 tan 2 0), —6 r 2 tan 5 sec* 0. 
6. (a) 2x, 2(x 4- tan x sec 2 x); 

p ( 

Section B, Page 2S6 

sec y -f &cV 


2. *(«§ — ajfo) -j- y(y8 — aate) 


/JA M *V, . dV f dV . BV\ dV 

(b) cos 0 h sin 0 — » r cos 0 sin 0 — ) > — • 

dx dy \ dy dx / dz 


1. (a) y' - - 

<»* 

2 


x sec y tan y 4 2x 8 y * 

Sx^ 3x 2 y 

cos * — 3* 2 * ^ ** cos z — 3a 2 


!. ? i ? ( I v5 i T?g-v 5)iy| 7 (» v ^+?s + *D ! 


B. du ■■ 2x dx 4 2y dy «• 2r dr. 

g*1/ g*V 

e. /„ - ~2 ~ — 2 + *»)>/« “ “=— -r 

r 4r uHr 


r — ux). 


Section 8, Page 246 

1. (*■ 4 1)/V5 

8 . H[S y/s 4 1 4 e(l 4 VS)] * 6.811. 

Section 9, Page 249 

1. a/3, a/3, a/3. 2. 8a6c/3 y/%. 8. a/3, 6/3, c/3. 

4. y/SP/( 2 VS 4 3), (VS 4 1)172(2 V3 4 3), P/( 2^ 4 3). 

g. I m h - ~ VoOr^F, d - V6 «. 


Section 10, Page 264 

8. (a) 106°46', 90°; (6) 164°16', 90.° 6. d/Va* + 6* + c*. 

Section 12, Pages 260-261 

i. j + (j - ft + («• - m + («• - 1 )kk + k* 

+ S*‘ + T*'‘ + ( ,+ i)"* + i‘ , + ' • ■ 

where & X — „1, k » y — - 
2 

8. « |l + (* + *) + [ft* + 4ta + fc»] + • • •}> fc - * - 1, it - V - ] 
*• 1 + x + (** - y 1 ) + i (i* - 3xy J ) + jj (* 4 - 6*V + y 4 ) + • • 

Section 18, Pages 268-264 

4 irsin (rar/2) t COS (rnar/2) — 1 A 

X * £ +~ ? * “** 


axm- 



786 


ANSWERS 


S.«g-l°g2). 

8. 2x*. 


4. — tan a. 

7. air (a* ~ 1)~* 


Section 14, Page 869 

d 

*• g (py') - n - S - o. 

Section 18, Paxes 379-277 

3 . u*vdudvdw. 

3. «a»(r/2 - %). 


1. 

8. 32o‘/9. 

Section 17, Page 881 

JL Tt?/2. 

4. 8a*. 


3. ududv. 

6. x(l - e _ °*). 


2. 4a*(x/2 - 1). 

5. t ~ a cos* (o/2). 


CHAPTER 4 

Section 2, Page 291 

2. A + B + C - 0. 8. A - i^(S + D), B - H(S - D). 

8. f°> TaJ ; <b) 2 (jaT ^ TbT )' 

Section 2, Page 298 

1. (a) 6j, — 5j; (6) A + B - 21 + 3j + 4k; (A + B) + C - 31 + 3j + 3k; 

B + C ■» 2i + j; associative law; 

(e) 51 + 10j X 15k, — 2i — 4j — 6k, 3i + 6j + 9k, 31 + 6j + 9k; 

(<J) 3i + 6j + 9k, 3i + 3j + 3k, 6i + 9j + 12k; 

(e) -4. 

Section 4, Page 294 

1. (a) 10, 2, 8; (6) 6, 4, i + 3j + k, 10; (c) 12; 

(' d) cos -1 3/V2I; (e) 4/>/5; (/) s - 4; (g) -i - j + k. 

2. (6) x — —20; y «■ 8; * — 1. 

Section 6, Pages 296-297 

1. (a) — 2i + 3j — 4k, 51 -_4j + 3k, 3i — j — k, 21 + 3j + 3k, 81 - j - k; 
(e) 131 + 2j + 2k; (d) y/l 77/2. 

Section 6, Pages 298-299 

2. (a) 0; (6) * - %; (d) 0, 0. 

Section 7, Pages 301-302 

1. (a) R'(t) - 21+ 6tj + 3t*k; ( b ) R'(l) - 21 + 6j + 3k; 

(c) v - 2i + 6 j + 3k, |v| - 7. 

8. (a) v - R'(<) - 1 + j cos t - k sin t; |R'(«)| - -y/5; 

(6) s - 2VS. 

Section 8, Pages 306-306 

1. (a) W - 0; (6) W - -2; (c) W - 4. 

2. (o) T - -i + 2j - 3k; (6) T - 2i + 4j + 4k. 



ANSWERS 


787 


8. («)▼■> AsA x B; (6) t •» kC X (A — B). 

8. (a) R - XV* + 5j); (6) R(l,2) - H(4i + j). 

Section 9, Page 808 

1. (a) n - 1 + 2j + 3k; (6) cos" 1 6/V32; (c) 9/VU. 

8. (a) i + 2j + 3k; GO * - 1 + <; y - 21; * - 1 + 3/; (e) V5?; 

(e) *£1 + j + (/) Ml - Hi + Hk. 

8. (a) R - 6i - 2j + (— 4i + 4j - k)t; (6) -41 + 4j - k; 

(d) -Ax + 4tf - * + c - 0; (e) R - -31 + k + (-41 + 4j - k )(. 
4. (a) & - 31 s + 8; (b) t - 0; D - 2\/2. 

8. R - (i + j + 3k)i, (-«, -*). 

Section 10, Page 811 

1. n « oi -f 6j + ck. 

2. (a) ~i + 6j 4- 2k; (c) 2! + j + 3k + (-i 4- 6j + 2k )t - R . 

4. 0 « cos" 1 9/VTO^. 

6, — 16i 4- 8j 4- 4k. 


Section 11, Pages 815-816 


1. («) (-1 + j - k)/V3; GO -* + (» + 2) — (* — 2) - 0; 

(c) -y/3 + log (1 + )• 

2. (o) v « 21i + 2j + 21k; A - 21 + 2k; (6) v - 2V / 21* + 1; 


/x V4< s + 2 „ 1 — 21 j + k 

w * ” om,2 , "re* ; N “ -rrrr-rr- 


2(2 i* + l) 1 ’ 

8. Let R(l) “* ( 02 I 1 + fljl + ao)i + (& 2 I 2 + bit + &o)j 4" (cal* 4* cil 4* co)k; then wa, 
equation of the plane through the plane curve is 

(bic 2 - Mi)(* ~ Oo) -f ( 02 C! - aic 2 )(y - 6q) *f (Mi ~ Ms)(s - 4>) - 0. 

8. (a) T - (1 4- 2j)/v/5; N - (j - 2i)/V5; 

(6) V = i + 2j; A - 2j; (c) F ( - V5; A, - 4/y'S; 

(d) s' - Vl +41* ; s" - 41/V 1 + 41*. 

8. (a) A n - 2/V5; (6) * - 2/(5\/5 ). 


CHAPTER 6 

Section 2, Pages 866-867 

2. 0n * 1, 022 “ p 2 , fits “ p 2 «in 2 0, 012 " 023 *■ 013 ** 0, where p «* %\\ 

0 « X 2 ; 0 « X,. 

Section 8, Page 872 

1. At (1,2,3), Vu * 2i -f 4j 4- 6k; du/dn - 2\/l4; 

At (0,1,2), Vu - 2j 4- 4k; du/dn - 2 VS. 

2 . (a) ~(ix 4- jy 4- kzKx 2 4- y 2 4- * 2 )“*; 

(6) 2(ix 4- jy 4- k*)0c 2 4- y 2 4* * 2 )~*. 

8. du/dn * -3/V6. 

6. n « HO -2J4- 2k). 

8. du/dn «• — 3. 

Section 4, Pages 877-878 


4. du/dn ~ -7/ VS- 

6. H(2i - 2j - k), H(-2* 4- 2j 4- k). 

9. dv/ds - 6/ VS. 


8 


3. (a) Hi W -X; («) 



788 


ANSWERS 


4. Helical path, t 2 /8 — 1; rectilinear path, ir 2 /8 — X. 

(a) W m -K (6) Wm-g. 

Section 5, Page 382 

1. %■ 2. u{x,y) - xy + Hv* - K*?- 

4. (o) « - xyz; (6) xyz. 5. 0; u - log r. 

6 , 0 , 

8. & +v* - (*? + v? + 4)- H . 

Section 7, Page 888 

1. (a) 3; (6) 2/r; (c) 0. 2. 0, 2/r. 5. Su, 0. 

Section 8, Pages 390-891 

2. iro’t. 8. 4ira6c. 4. 4 ra 5 . 

Section 9, Pages 395-896 


1. (o) v"374; («) i%. 


a. (a) ▼ 

(&) v 


»(** + V s ) + j2®v; u - **/3 + iv 1 ; 

1 - V* ■ , 2/ ■ 1 y 2 - 1 

(!+*)* (1+ x )* J ' “ “ 2 (1+ *)* : 


(c) y •> iy cos x 4* j sin x; u — y mn x; 

(d) v - »*v(l - z 1 )-* - j(l - z*) H ; « - — v(l - z 1 )*; 

(e) v - i(z + 1) + j(v + 1); u - M[(* + l) 1 + (v + l) 1 ]. 

8. (o) 2ir; (6) 2 t. 4. 6ir. 8. (J r , 


Section 10, Page 899 

2. (a) 0; (ft) 0; (c) 0. 


Section 11, Page 402 

2. —t. 3 . 0. 4. o. 

Section 12, Page 406 

2* w « j (xy - Hs?) 4- k[*(* + y) - K(* 2 4- y 2 )J. 

8. -u — xy 2 4* x 2 * 2 — x. 

4. w « j(xy* 2 - Hxh) 4 - k(- + *V). 

6. No. 


Section 18, Page 408 

du . •idu du 
1. Vu-r l - + 7 - + k-. 

a. v 1 - - — + ■ 1 3 | L_. a * ■ 1 ** , l a* 

(u s + t^)u du (it 2 + t?)v dv u 1 + t?dv? li S + B 8 *! 1 + uVa^. 1 ' 
8. div F *■ — 3p cos 0/r 4 , curl P « 0. 

Section 16, Page 414 


2. Irrotational. 4. * - z 1 - y*, hyperbolas. 

8. Irrotational and solenoidal. 10. — 3z*y; * - z(z J + 



ANSWERS 


789 


CHAPTER 6 

Section 1, Page* 429-431 

2. (a) [1 + (x - oO 9 ]" 1 , -2(z - ofl[l + (x - erf)*]-*; 

(b) 2(1 + a¥)'\ 2(1 + **)->, (P - 2i + 2) -1 + (P + 2i + 2)"*. 

3. U m U — U K Uy m 0. 

7. (a) fiiv + »!*), /*(» + mtt). 

9. (o) Pi(y - ax) + Pj(y 4- ai); (b) Fi(y - 2x) + F t {y + x); 

(c) Fi(* + iy) + F j(x — *'y); (d) F i(y + x) 4- xF»(y + *). 

10. (d) Ft(y - fix) + F 2 (y + x) - y */ 60 + x«/6. 

11. (a) ~ + Fi(2y -f x) -f F*t(y — x); (6) ~ + Fi(y ax) -f F 2 (y -f ax); 

(c) 4- K 2 I/* + Fiiv - 2s) + F 2 ( 1 / - x). 


Section 4, Page 440 

2. u(0,ir) *» sin 2ar; t^(0,ir) «** ^ cos 2ax. 

a!(1 

3. a «* =fc — — H — — - > n «■ 0, dbl, =fc2, .... 

15 5 

4. Const + (2a)~ 1 e~’ t . 6. ae~*( 2x* - 1). 

Section 5, Pages 445-446 

1. - 1 , ±21, ±31, .... 


Section 6 , Page 449 

2 . (b) 6/2. 

Section 8, Pages 464-465 

1. 0.45 oscillation per sec. 
Section 9, Pages 458-459 

1. 2.07 X 10 6 cal /(m 2 ) (day). 
4. (c) ^4 1 1 ~ e“ <a " /i),< ]. 


3 . («)®E <- 

* n-0 


1 )" 


(2n + 1)* 


^ (2n 4* 1 )tx 
l 


Section 10, Pages 462-463 

4. (6) u(x,t) — 2c n e“ Ia(2n ~ 1)/2QS *sin7 (2n 


l 


l)x; 


2 r l 

c« « T / fix) sin - (2 n — l)x dx. 
t Jo l 

;/>>« 

/: 


_ „ nrat . nrx 

7. Za n cos — sm ~J~ * °n ■ 

a vi « nir <^ , nrx 

8. 2*6» sm — sm ; 6 W « 

l l 

Section 11, Pages 466-467 

3. 0.44883, 0.14922, 0.00004. 

r , x 400 

6. u(z,y) m — 2J 


nrx , 
sin -y~ ax. 


2 

nra 


0(x) sin dx. 


— i— - e -o«-i)»*/io gin (2n _ d If. 

2n - 1 ' to 



790 


ANSWERS 


6. u(x,t) “ ^ 2n ~ — j sin (2n — 1)^2- 

7 . 36 . 5 , 41 . 9 . 

Section IS, Page 471 

n 

^ 2t Jo r 2 — 2r# cos (0 — <*>) -f # 2 


i2 2 - r* 


i* 2 -r 2 


o L# 2 — 2#rcos(0 — «*>) 4* r 2 R* — 2i2r cos (0 + <*>) -f r 2 
2t r 2 - i? 2 

Md*. 


]/(♦)<*. 


Section 13, Page 474 

200 , 

1 . U - 50 + — , 


sin (2n — 1)0. 


d<f>. 


(2 n - l)** 2 ”- 1 

60 f T a 2 - r 2 

2 . U ** — / 

sr Jo a 2 ~ 2ar cos (0 — <£) -f- r 2 

*• U “ V Z (2n~l)a^ r2 " -1 8m (2n - 1)9 ' 


4. u(x,y ) * 2an sin r?u sinh iray; a n =* -7*“ — / /(s) sin *nx dx. 

smh %n Jo 


Section 14, Page 479 

l\ * 


Section 16, Page 482 

1. £ Jo(frnO COS m ^ Z(A mn COS « mn * + Bmn Hm O&n - 7* 

2. JZA n e~ a!lk n t Jo(k n r) t where 1 = 2A n Jo(k n r). 

Section 18, Page 490 

a. («A)-* (W— > cosh <a. 

Jo 2ort 

S. (iraH)~ H f fU)e- ixUP>Kia, ‘ ) ainh~dt 
Jo 2ort 

4. (4«A)~ H jf /(*) j £ [e~ (l_2 ’ li_,),/<4a ’ 1) ± e -(*-l» , +.)’/(4o s i)]| dj _ 


Section 19, Page 493 

3. u(*,j/,0 - (4* ^at)- 1 f f e -[(*-» 1 ) , +(v-n) , l/U« , iy( ;Cl)l/l ) ^ dyi 

j — <30 J-~<X> 

4. u(x, »,*,() = ni7-Tii5 


fivuh) 


6.»(x.„,0 -fj_" W(t _ (i) 

«(*,<)- fjI" Wr mH dti. 

</o [4ira 2 (t - ti)J^ 





ANSWEBS 


m 


Section 24, Page 607 

1. |fc|<2, ~2, >2. 

2* Respectively m, in or outside the unit circle x* 4* y* *■ 1. 

5. (a) Ice v + (1 - c)e~*] sin x. 

Section 26 t Page 512 

2, At the lattice points in the region y > x t x > Q; y > h — & f z <i 0. 

S. 


Section 27, Page 514 
1. 19. 


2 . 2 . 


S. 2. 


Section 28, Page 517 
12 
' RL 


1. 7(x,Q - 


MCUA). 


-(llROinrfiyU *** , 


<50 ir n-»l n 100 

». / - 0.6 + 1.1 £ (-1)" COB /0 ‘. 

»-i 1,000 


Section 29, Page 519 

3. t/„ - 0; Urr - 0; l/rr + ^ - 0. 


CHAPTER 7 

Section 1, Page 588 

1. (o) 2, x/3; (5) 2>/2, */4; (c) 2, *; (d) 1, 3ir/2; 

(e) V2A 7ir/4; (/) 1, */2; (y) fc, r/3; (A) 4, 

2. (a) —8; (6) -1 + 1 ; (c) (2 - V5 )/2 - /(2 + V3 )/2. 

8. (a) 1} (b) 1; (c) 1. 

4. cob <v/6) 4* * sin (r/6), cos (x/6 -f 2 t/ 3) -f i sin (r/6 -f 2r/3), 
cob (ir/6 + 4ir/3) + i flin (ir/6 -f* 4x/3). 

7. (a) 1, JS<-1 + ), Hi -1 ~ tV3 ); (6) 1, t, -i, -1. 

16. (a) Circle x 2 4- y 2 * 1 ; (&) circular region x 2 -f y 2 < 1 ; 

(c) region exterior to the circle x 2 4- y 2 1 including the boundary. 

18. (a) Circle radius 2, center at (1,0); ( b ) circle radius l/\/const, center at (0,0). 

Section 2, Page 585 

1. (a) (x 2 - y* - X -f 1) 4* »(2xy - y); (6) x/(x* + y 2 ) - iy(x* 4- y 1 ); 

(d) (x 2 -f y 2 - i)/[ar» 4~ (y 4- 1) 2 ] - ^/[x 2 4- (y 4- l) 2 ]; 

(/) x + t'2y; (g) (x 2 4- y 2 )“ l . 

2. (a) Open region x<3, — »<y<<»; 

(6) The region y>l, — »<«<*>; 

(c) The region exterior to the circle of radius 1 with center at the origin and in- 
cluding circular boundary; 

(d) Circular ring centered at the origin with interior radius 1, exterior radius % 
including the boundary of the inner circle; 

(«) Open circular region with center at (1,0) of radius 1; 



792 


ANSWBBS 


if) Closed circular region of radius 1 with center at *o ■" *o 4 tjfo; 

(g) Open region exterior to the circle of radius 2 with center at (0, — 1). 

Section 8, Pages 989-540 

*. («) M(« _1 + *) coe 2 + - e) ein 2; (c) Io « '''»-(./<+«*.); („) «wi+»r 

4. (c) J$(« + « _J ) dn 1 + H*’(« — « -1 ) co» 1; 

(d) «*^**(ooe 2*y + % sin 2a?y); 

(e) e* /< ^ + ^ ) {cos [!//(«* 4 y*)J - i sin {y/(a^ 4 y*)]|. 

5. (o) log 4 4- (W log 5 4 i-ar/2; (e) e~ T/a ; (/) «(cos 1 4 i sin 1); 

(ff) HU* - «-*). 

6. (a)(r 4 2dfc)i; (i>) r/2 4 2r* -ilog(2db y/S); (d)2*k,k - 0, 41, db2, 

Section 4, Page 545 

8 . Q>) % «■ —1; (c) i «• 0; (d) * — r/2 4 fcir, fc « 0, =bl, ±2, . . .; (e) * — 1, t » — 1 

(/)» (y)i W at all points; (i) * «■ 0, (k) z « d=i. 

Section 5, Page 548 

1. M(2 4 lli). 

2. 1 along rectilinear, 1 4 i/3 along parabolic. 

5, 0. 6. 0. 7. (—2). . 

Section 7, Page 555 

2. 2wt. 3. 2, upper half; —2, lower half. 

5. 0. 10. (a) 0; (6) 0; (c) -*»; (d) ri. 

Section 8, Page 559 

1. 2*i(8 - 13i). 8 . (a) 0; (b) 2 t i; (d) 0; (e) 2ri. 

6. ~2*i. 


Section 9, Page 561 

1. u ^ x* — &ry 2 . 

2. (o) z 4 iy\ (b) cosh y cos x — i sinh y sin z; (d) e z (coa y 4 i sin y). 


Section 10, Pages 564-565 

a .f 

nZl n 

oo -n « / i\«- 1 2 2»—1 

*• (0) S »!' * “ (fc) .?x (2n — 1)1 * “ * : 

(d) E (-1)— «-l. 

«— 1 ft 


Section 11, Page 569 


1. (o)i + 2 + 3* +4^+ •••; 
z 

<&) (nb)» + i~i + 1 + - *) + o - *) 3 + a - *)‘ + • • •• 

„ . _i , j_ j_ , j__ 

*» + 2!*< “ 31»* ^ 4!*» 



788 


ANSWSRS 


S. Functions are expressed in Laurent's series. 

4. (a) - — i- - l _ (* - 1) - - 1)* ; 

Z —■* x 


2 3 1+2* 1+2* 

< 6) ; + ? + -£- + -^- + 
i * ** «* . i 


^ 2 2* ~ 2 s 2 4 


+ -+-, + -! + 


Section 12, Pages 672-673 


( 6 ) 1 --,+ 


2^ 

2! z 4 


I i 

3!*< 


4 + 


, residue 
, residue (0); 


y v 1 11 z z 2 . . , 

w r* _ ; + 2i _ 3! + 5i~" , ’ res,due( " 1); 

(/) « 2 +* + £j +™j + •••, residue (HO; 


(if) 


2 4 8_ 

z* 2!z 2 3 Iz 

( i ) Residue — at z * 
6. No. 


16 32s 

4 ! 5 ! 

1, residue }$ at- z » 


- •••, residue (~%Q: 

1. 


Section 13, Page 674 

1. (a) ~ri; (c) 2irt/3!; (d) -8«/3; (/) 0°. 

2. 2Tt. 4. (a) 0; (b) 2ri. 

Section 18, Page 586 

1. (o) cos x cosh y, sin x sinh y; (6) e x cos y, e x sin y; 

(d) log ( x 5 + y 1 )*, tan -1 (y/x); (e) x/(x* + y 1 ), -y/iz 2 + y 2 ). 

2. (c) v « e* sin y — x; (d) 8inh z sin y. 


CHAPTER 8 


Section 1, Page 612 

1. H. Hs. 

». H. 

6. 81, 71, 71/2. 

7. (a) X; (b) 0; (c) Ms; W H- 

Sectioo 2, Page 616 

l. W, H, H, X, X, H, X . X- 

6. Mi- 

Section 8, Pages 621-622 

1. 33/16,660. 

•• Xi, Xl, Xt- 


2- Ho, X , Ho, Xs, 1,323/46,189. 
4815 ! 13147 ! 47151 

‘ 52! ’ 8152! ’ 30 52! ' 

12 2m + 2n — 4 

6 . —t — 

n m mn 


2. Questions 1, 2, 8 can be answered. 

7. X- 


2. %, Ha. 

4. ‘Ho- 



794 


ANSWERS 


». H, H». 

7 . Ho, Ho- 

51 5! 18,781 

52* < 52 . . . 48’ 13‘ 


6. » > (log 2)/(log fl — log 5). 
„ /3 \* 1211 ... 8 

‘■U > 


52-51 


48 


270,840 
52-51 -50-49* 


10- 1 - j>* n - 3. 
Section 4, Pages 026-627 


1- 8 p. 2. E(XY) ** 6 pq. 

8. (a) 1; (6) 1; (c) 1. 4. X 2 + Hi ~f Xo + * * • + H 

6 - 2 . 


Section 6, Page 681 


1. Hit Hit l %2 1 l %2t Hi, X*. 4. 


Section 6, Page 687 

1. 0.75, 0,60, F(x) - x 8 for 0 < x < 1, m - 0.794. 

2. Xt, 0.206. 8. 

4. H - 6- X- 

Section 7, Pages 640-641 

1. W«, H*> H, 2 and 3. 2. (a) 125/3,888, (6) 2% 48 . 

8. (0.65) 10 + 10(0.65) 9 (0.35) + 45(0.65) 8 (0.35) 2 *f 120(0.65) 7 (0.35)». 

4. 5(X>« + 4 (H) b (H) + 45(X) 4 (X) 2 + 40(X) 8 (X) 8 + 15(X) 2 (X) 4 . 

B. 741/2,728. 7. 0.57, 0.57, - - — 

n n 


Section 8, Page 644 

1. 0.499. 

Section 9, Page 660 

1- 0-039. 

8. 0.979. 

5- 0.083, 0.166, no. 

Section 10, Page 664 

1. 1,540. 

3. *X 0 . 

Section 11, Page 668 

1. (a) 0,368, 0.402; (b) 0.368, 0.373. 
8 - 1,005. 

6. 0.577. 


2. *(1.5) - *(-1.0) * 0.806. 
4. 0.0222. 


2. 46,413/78,125. 
4. 0.91854. 


2. 0.758. 

4. Expected number » 5. 


Section 12, Page 668 
2. 0.82. 


Section 18, Page 667 

1. 

8 - X V5 - 0.577. 


6.83. 


2 . H, H, *H,ye*,P - -l. 



ANSWERS 


70 $ 


Section Iff, Pages 070-671 

1. 7.5, 0.26. 2. 0.145, 54. 

CHAPTER 9 

Section 1, Page 679 

1. (a) -0.8 < x x < -0.7, a* - 2, x 9 - 4; 

(6) -0.8 < %i < -0.7, 1.2 < 3 2 < 1.3; 

(c) -0.8 <xi< -0.7, -0.6 < 32 < -0.5, 1.0 < 3* < 1.1; 

(d) -0.6 < xi < 0.5; (e) 4.4 < 3i < 4.5. 

2. h ~ 1.23. 

Section 2, Page 684 

1. Prob. 1: (a) -0.75; (6) -0.73, 1.22; (c) -0.77, -0.55, 1.08; ( d ) -0.57; (e) 4.49: 
Prob. 2: 1.226. 

2. -0.942, -0.200, 1.045. 

Section 3, Page 686 

3. 2.310 radians. 

4. (a) 0.739; (6) 0.667; (c) -0.725, 1.221; (d) 2.924; (e) 1.045, -0.942, -0.200. 
Section 4, Pages 688-689 

(a) x « x Hb; y - 2 Hb', * ~ -Ha'p 

(b) x\ « 1; 3a « — 1; x% «• —2; 34 — 3; 

(c) xi ~ -0.107; 32 - 0.988; x z - 0.317. 

Section 7, Pages 694-696 

1. y « 0.253* - 0.503 -f 0.25. 

Section 8, Page 696 

2. 9.466, 12.549. 

Section 9, Page 700 

1. 2.784, 2.700. 2. If % « 60, 9 - 40.82, 42.52, 42.50. 

3. 2.581, 2.627. 4. 106.09. 

Section 10, Page 702 

1. (a) y - K* + H; (5) y - 2.53° *; (c) 0.3(10° **). 

Section 11, Page 711 

1. y - 4.98 - 3.13x + 1.263 2 . 

3. It - 1.778; 6 - 1.9349; 9 « 60.02(0.861)*. 

Section 12, Page 715 

1 . y ** 0.75 4* 0.10 cos x — 0.05 cos 3x — 0.29 sin x. 

2. y - 0.85 - 0.25 cos 2x - 0.05 cos 4x + 0.05 cos 6x + 0.26 sin 2x - 0.03 sin 4*. 

Section 13, Pages 720-721 

1. 25.252, 25.068. 

8. 128.6. 

6. 39.30, 38.98. 


2. 132.137. 

4. 666.25, 666.00. 



ANSWERS 


706 

Section 14, Page 738 

1. y(±0.2) - 1; y(±0.4) - 1.02; y(±0.6) - 1.061; y(±0.8) - 1.124; y(±1.0) 
*• 1.214. The corresponding exact values are 1.010, 1.041, 1.094, 1.174, 1.284. 

*. v - vi + y'(ii)(* - *0 + — (v'ta) - v'fa>)K* - si) 1 - 
2 h 

8. yi - 1.0100; j/j - 1.0403; y t - 1.0927. 

Section IS, Pages 726-727 

1. yt - 2.0442; y 1 - 2.3274; yt - 2.6509; y» - 3.0190; y w - 3.4363. 

3. y( 0.3) - 1.3498; y(0.4) - 1.4917; y(0.5) - 1.6485; y( 0.6) - 1.8218. 

4. 0.2740. 

Section 16, Page 780 

1. y « x 4 - x 2 + M** 4- H 4- Ra; * - x 4- M** + 

2. 0.1, 0.2205, 0.3627, 0.5281. 

8 . 1.01, 1.031, 1.063. 

4 . v l + + M*? + H2X 4 + Ho** 

z m X + 4- H® 3 4* H 2^ 4 4 * • 

5. 1.0052, 1.0215, 1.0502. 


APPENDIX A 

Section 1, Page 747 

1. (18, 0, -1, 0). 2. (18, 0, -1, 0). 8. 1. 

Section 2, Pages 752-768 

1. (a) (2, -1, 1); (6) (1, «, -Jf); (c) (3, -1, 2); (d) (1, r l, -2, 3). 

2. (a) (~*/7, S*/7, A?); (5) (0, 0); (c) (0, 0, 0); (d) (l fc/4 , 7*/8, *); W ft 2k, 0); 
(/) (0, 0, 0). 

8. (a) (1, —1); (6) inconsistent; (c) inconsistent; ( d ) (1, 3 k —2, &). 


APPENDIX B 

Section 2, Pages 758-769 

2. y * coe 2x 4- H sin x — H »in 2x. 8. y - 3^e x — 4- M** — x. 

6.if-24«~ 2 *- xe"~® - 3c"~ x . 6. y - e 

Section 8, Pages 761-762 

1 * V *• fSc"** - 2. y «* 1, y *» x, y - for x > 0. 

Section 4, Pages 764-765 

1 . \ e“' 2x cosh -7== x — - \/3 «~ 2x sinh x. 

3 \/3 2 V% 

SL (a) cos x; (b) e~ x cosh x 4 - Mt~* sink x ** ?£ 4 - K«~ 2as ; 

(c) e““ H *(co8h H\/6 x — sinh 4£\/6 x). 

Section 5, Page 767 
tie 

J 0 ^0 





ANSWERS 


797 


a. (•) J4 - «*■* + (b) K«e** - ««- + 

(c) Hx-H+t-*- He-**. 

Section 6, Pages 768-789 

+ ( t cuL — g -(»/A)< _ oai cos u( + cR sin at with c *» 

l e -(RIL) t 


fi* + «*£*' 


S. L~ l f V(r)«-< B /«<‘“ T )dr - R - 1 fV( t )[1 - *—€»/«<*— r>) (Jr 

Jo JO 


+ m u _^, 





INDEX 


The letter p. after a page number refers 

Abel's theorem, on differential equations, 
54 p. 

on power series, 142 
Absolute convergence, of integrals, 755 
of series, 127, 170 

Absolute value of complex numbers, 168, 
528 

Acceleration, normal, 313 
tangential, 313 
vector, 302 p. t 313 
in cylindrical coordinates, 367p. 
Adams' method, 723, 728 
Addition, of complex numbers, 166, 529 
of matrices, 327 
parallelogram law of, 288, 529 
of series, 117 
of vectors, 288, 317 
Adiabatic expansion, 50 p. 

Algebraic equations, 677 
solution of, by graphical methods, 678 
by iterative methods, 679 
systems of linear, 350, 687, 689, 749 
Alternating series, 128 
Amplitude, of simple harmonic motion, 44 
of waves, 428 
Analytic functions, 540 
branch points of, 570 
Cauchy's formula for, 555 
Cauchy's theorem for, 547 
differentiation of, 541, 557 
essential singular points of, 570, 574 
geometric representation of, 575 
integrals of, 545, 547, 551 
Laurent's expansion for, 565 
mapping by, 575-594 
maximum modulus theorem for, 558, 
561 

poles of, 570 
residue theorem for, 573 


to a problem, the letter n. to a footnote. 

Analytic functions, residues of, 570 
singular points of, 543, 569 
Taylor's series for, 561 
Angle, phase, 44, 528 
solid, 399p. 

Angular momentum, 305 
Angular velocity, 302, 399p. 

Antenna, radiation from, 486 
Arc length, 301 
in curvilinear coordinates, 362 
of an ellipse, 147 
Argand’s diagram, 528 
Argument of a complex number, 528 
Arithmetic means, 667 
Asymptotic equality, 12, 123 
Atmospheric pressure, 50 
(See also Pressure) 

Attraction, of a cone, 277 p. 

Coulomb’s law of, 408n. 
of a cylinder, 277p. 

Newton's law of, 46 
of a sphere, 47, 277 p., 410 
Augmented matrix, 750 
Average, arithmetic, 667 
Average-value theorem, 498 


Base or coordinate vectors, 319 
in curvilinear coordinates, 363 
in cylindrical coordinates, 367p. 
orthonormal, 321 
in spherical coordinates, 367p. 
transformation of, 337, 367p. 

Basis, 319 

Beams, bending of, 15 
buckling of, 95p. 
cantilever, 16 

on elastic foundations, 86p., 94p. 
vibration of, 435p. 


799 



800 


INDEX 


Bending moment, 16, 435p. 

Bernoulli's differential equation, 27 
Bernoulli-Euler law, 16 
Bessel's differential equation, 159 
Bessel’s functions, 162, 198, 480 
asymptotic formulas for, 199p 
expansion in series of, 481 
generating function for, 166p. 
orthogonality of, 198 
zeros of, 198 
Bessel’s inequality, 202 
Beta function, 149, 765 
Biharmonic equation, 430p. 

Bilinear forms, 349 
Bilinear transformation, 577 p. 

Binomial distribution, 639 
generating function for, 6 4 Op. 
Binomial frequency function, 639 
Binomial law of probability, 639 
Laplace’s approximation to, 647 
normal approximation to, 647 
Binomial theorem, 155 
Binormal, 312 
Boltzmann constant, 633 
Boundary-value problems, 91, 442, 730 
Bounds for Fourier coefficients, 211 
Branch points, 570 
Buffon’s needle problem, 637 


Cable, flow of electricity in, 514 
hanging under gravity, 40, 454 
oscillations of, 445, 454 
supporting roadway, 42p. 

Calculus of variations, 264 
isoperimetric problems, 269 
problems with constraints, 269 
Cantilever beam, 16 
Cartesian reference frames, 321 
Catenary, 41 

Cauchy’s convergence criterion, 115 
Cauchy’s differential equation, 78 
Cauchy’s inequality, 322 
Cauchy’s integral formula, 555 
Cauchy’s integral test, 120 
Cauchy’s integral theorem, 547 
Cauchy’s principal value of an integral, 
602 

Cauchy-Riemann equations, 413 
Oauchy-Schwarz inequality, 322 
Center, of gravity, 275, 281 


Center, of mass, 44, 303 
motion of, 44, 304 
Chain under gravity, 40 
Chain rule, 228 

Change of variables, in functions, 237 
in integrals, 270 
Channel, flow from, 594 
Chaplygin’s method, 37 
Characteristic equation, 54, 67, 521 , 733 
of a matrix, 344 

for systems of linear differential equa- 
tions, 100, 106p., 733 
Characteristic frequencies, 477, 479p., 481, 
482p. 

Characteristic functions, 732 
Characteristic values, 344, 507p., 732 
Characteristic vectors, 344 
Characteristics, 440, 508, 517 
discontinuities on, 519 
Chemical combinations, 14 
Chemical reactions, 15 
Circle of convergence, 170, 562 
Circulation, 397, 591 
Clairaut’s equation, 27 p. 

Cofactors, 741 

Column, axially loaded, 86p., 90 
Euler's critical load for, 92 
Combinations, 611 
Combinatory analysis, 611 
Comparison tests, for integrals, 755n. 

for series, 122, 125, 134 
Complementary function, 59 
Complex function, 534 
continuity of, 540 
differentiation of, 541 
integration of, 543 
Complex numbers, 166, 527 
absolute value of, 168, 528 
addition of, 166, 529 
argument of, principal, 528 
conjugate, 529 
modulus of, 528 
operations on, 528 
phase angle of, 528 
polar form of, 528 
roots of, 531 
Complex potential, 587 
Complex roots of unity, 532 
Complex variable, elementary function* 
of, 534 

Complex-variable theory, 523-604 
Components of a vector, 291, 317, 321 



INDEX 


801 


Composite functions, 280 
Compound probability, 617, 635 
Condenser discharge, 81, 84 
Conductivity, thermal, 415, 455 
Conformal mapping, 583, 598 
examples of, 575-595 
invariance of harmonic functions under, 
585 

Riemann’s theorem on, 586 
Schwarx-Christoffel formula for, 599 
Conjugate complex numbers, 529 
Conjugate harmonic functions, 560, 588 
Conjugate matrix, 343 
Conservation, of energy, 43 
of matter, 417 
of momentum, 44 
Conservative force fields, 408 
Constraints in calculus of variations, 269 
Continuity, 218 
of complex functions, 540 
equation of, 412, 417 
piecewise, 755 
of scalar functions, 368 
of vector functions, 299, 368 
Contour integrals (see Line integrals) 
Convergence, circle of, 170, 562 
interval of, 139 
radius of, 139, 170, 562 
of series ( see Series) 
uniform, 132 
Convolution, 488 
Convolution theorem, 488, 762 
Coordinate lines, 272, 359 
Coordinate surfaces, 359 
Coordinate vectors, 319 
Coordinates, affine, 366 
curvilinear, 357 
divergence in, 406 
gradient in, 407 
volume in, 364 

cylindrical ( see Cylindrical coordinates) 
orthogonal, 363 
parabolic, 408p. 
spherical, 360, 367p. 

Correlation coefficient, 666 
Coulomb's law, 408n., 467 
Couple, 305 
Covariance, 664 
Cramer's rule, 326, 749 
Cross product, 294 
Grout's reduction, 687». 

Curl, 396 


Curl, in cartesian coordinates, 398 
in curvilinear coordinates, 406 
relation to rotation, 399p. 

Current flow, 416 
in cables, 514 

in electrical circuits, 81, 87, 100, 76 lp., 
768p. 

Curvature, 150, 311 
Curve, elastic, 16, 86p. 

Frenet’s formulas for, 311 
integral, 7 
length of, 301 
minimising, 265 

of minimum descent, 269p., 767 
motion on, 301, 313 
normal to, 311 

piecewise or sectionally smooth, 372 
pursuit, 33 
on a surface, 309 
trihedral associated with, 312 
Curve fitting, by finite differences, 694 
by graphical means, 701 
by least squares, 702 
by trigonometric functions, 711 
Curvilinear coordinates (see Coordinates) 
Cycloid, motion on, 46 p., 269p., 767 
Cylindrical coordinates, 359 
acceleration components in, 367p. 
base vectors in, 367p. 
velocity components in, 367p. 
volume element in, 365 


D, 57 
V, 370 
V 2 , 387, 407 
A, 510, 692 
5, 761 

D'Alembert’s solution of wave equation, 
439, 485 

Damped oscillations, 449 
Damping, viscous, 82, 449 
Definite integrals (see Integrals) 
Deformation of contours, 549 
Del, V (see Gradient) 
de Moivre's formula, 530 
Dependence, linear, 52, 70, 317 
Derivative, directional, 243, 253, 369 
(See also Gradient) 
normal, 244, 253, 369 
partial, 219 

Determinants, 325, 741-758 



Determinants, cof actors of, 741 
differentiation of, 743 
expansion of, 326, 741 
minors of, 741 
multiplication of, 325, 745 
solution of equations by, 326, 748 
Wronskian, 52, 54 p., 71 
Difference equations, 510, 734 
Dirichlet’s problem for, 511 
elliptic, 511, 518 
hyperbolic, 513, 518 
parabolic, 512, 518 
Difference operators, 510, 692 
Differences, backward, 692 
finite, method of, 734 
forward, 691 
Differential, 223, 310 
approximations by, 226, 311 
of arc length, 362 
exact, 226, 380 
total, 226, 234 
of volume, 364 

Differential equations, elliptic, 505, 511, 
518 

Euler’s, 78, 267 
exact, 20 
hyperbolic, 507 
Lagrange’s, 26 

ordinary (see Ordinary differential equa- 
tions) 

parabolic, 506, 512 

partial (see Partial differential equa- 
tions) 

systems of, 95, 733 
Differential form, quadratic, 362 
Differential operators, 57, 430p. 
Differentiability, 226 

Differentiation, of analytic functions, 542 
chain rule for, 228 
of composite functions, 230 
of definite integrals, 261 
of determinants, 743 
of Fourier series, 210 
of implicit functions, 230, 235 
of infinite series, 135 
numerical, 698 
partial, 219 
of power series, 140 
of vector functions, 299 
' Diffusion, 14, 416, 463n. 

Difftiflivity, 463n. 

Dimensional analysis, 433 


Dipole, 408p., 496 
Dirac's delta function, 761 
Dirac's distribution, 759 
Direction cosines, 370 
Directional derivative, 243, 253, 369 
(See also Gradient) 

Dirichlet’s conditions, 180 
Dirichlet’s kernel, 205 
Dirichlet’s problem, 467, 502, 611, 595 
for arbitrary regions, 595 
for a circle, 469 
for a half plane, 484 
for a half space, 503 
Dirichlet's theorem, 180 
Discontinuity, simple, 178 
Discrete distributions, 627, 628 
Discrete variables, 628 
Dispersion, 427 
Distribution function, 632 
Distributions, binomial, 639, 640p. 
bivariate, 666 
continuous, 631 
discrete, 627, 628 
Gaussian, 633 
Maxwell-Boltzmann's, 633 
normal, 633, 651 
Poisson’s, 633 
Divergence, 384 
in cartesian coordinates, 386 
in curvilinear coordinates, 406 
Divergence theorem, 388, 493 
Dot product (see Scalar product) 
Double layer, 497 
Dummy or summation index, 324 
Dynamics, laws of, 302 


e, 655n. 
e 4 *, 173, 536 
e*, 172, 536, 581 

Eigenfunction (characteristic function) 
732 

Eigenvalue (characteristic value), 344 
5Q7p., 732 

Eigenvector (characteristic vector), 344 
Elastic curve, 16, 86p. 

curvature of, 16 
Elasticity, 599 

Electric circuits, 81, 87, 100, 756 
Electromechanical analogies, 81 
Electron, acceleration of, 45, 46p. 
mass-to-charge ratio, lOOp. 



INDEX 


Electrostatic field, 592, 594 
Electrostatics, 457, 496 
Ellipse, length of, 147 
Elliptic differential equation, 505 
difference equation for, 511, 518 
Elliptic integrals, 49, 86p., 148 
Emissivity, 463 
Empirical formulas, 701 
Energy, conservation of, 43, 48 
kinetic, 43, 306p. 
potential, 43, 264 
Envelope, 35 

Equation of continuity, 412, 417 
Error function (probability integral), 
table, 776 

Errors, estimate of, 658, 660 
Gauss’ law of, 662 
mean-absolute, 662 
mean-square, 662 
probable, 662 

in solving differential equations, 38, 
104 

theory of, 658 

Essential singular points, 570, 574 
Estimate, of errors, 658, 660 
maximum likelihood, 640, 660 
reliability of, 670 
unbiased, 640 
of variance, 669 
Euclidean space, 321, 374n. 

Euler's critical load, 92 
Euler’s differential equation, 78 
invariational calculus, 267 
Euler’s formula, for exponentials, 173, 536 
for Fourier coefficients, 175, 196 
Euler’s hydrodynamical equations, 419 
Euler's polygonal curves, 721, 727 
Euler's theorem on homogeneous func- 
tions, 234 

Euler-Fourier formula, 175, 196 
Even functions, 183 
Fourier expansion for, 184 
Events in probability, 610, 618, 619, 638 
Exact differential, 226, 380 
Exact differential equations, 20 
Expansion, adiabatic, 50p. 
of determinants, 326, 741 
Fourier, 175, 196 
Heaviside, 766 
Laurent, 564, 565 
Maclaurin, 144 
in power series, 144 


Expansion, in series of orthogonal funo* 
tions, 201 
Taylor, 144 
Expectation, 623, 634 
of product, 663 
of sum, 624 

Expected frequency, 627 
Expected value, 623, 639 
Exponential function, 172, 536, 581 
Extrapolation formulas, 696 
Extreme values, 250, 264 
Extremum, 250 


Factor, integrating, 22 
Factorial, n !, approximation for, 644 
Factorial function, 162, 755 
(See also Gamma function) 

Falling bodies, 47 
Fermat’s principle, 264 
Field, 367 
conservative, 408 
electrostatic, 467, 496, 592, 594 
gravitational, 409, 467 
irrotational, 402 
solenoidal, 402 
Field theory, 355-420 
Finite differences, method of, 734 
Flexural rigidity, 93 
Fluid flow, 411, 416, 587-595 
under dam, 593 
ideal, 419, 587, 592 
incompressible, 412, 418 
irrotational, 412, 588 
out of channel, 594 
solenoidal, 412 
stagnation points in, 590 
steady, 412, 587 
vortex in, 591 
Flux, 384 
Force field, 408 

electrostatic, 467, 496, 592, 594 
gravitational, 409 
Forced vibrations, 86, 451 
Fourier coefficients, 175, 196 
bounds for, 211 

Parseval's equality for, 202, 204p. 
Fourier expansion, 175, 196 
for odd functions, 185 
Fourier heat equation, 414 
Fourier integral equation, 192 
Fourier integrals, 190, 194 



804 


INDEX 


Fourier series, 175, 196 
complex form of, 192 
convergence of, 200, 204 
differentiation of, 210 
doable, 476 

for even and odd functions, 184 
extension of interval for, 187 
integration of, 207 
uniqueness theorem for, 186 
Fourier transform, 194, 482-490 
Free vibrations, 79, 432, 444, 446, 475 
Frenet-Serret formulas, 312 
Fresnel integrals, 147, 153p. 

Frequency, characteristic, 477, 479p., 
482p. 

relative, 615, 638, 642 
resonant, 89, 477 
Frequency equation, 478, 481 
Frequency function, 627 
binomial, 639 
Fuchs’ theorem, 157 

Fundamental theorem of integral calculus, 
9, 261, 550 


Gamma function, 149p., 162 
Gas, ideal, 221 
viscosity of, 451 
Gauss’ distribution, 633 
Gauss’ divergence theorem, 388 
Gauss’ law of errors, 662 
Gauss’ reduction method, 350n., 687 
Gauss-Jordan reduction, 687n. 
Gauss-Seidel method, 689n. 
Geometric series, 115 
Gradient, V, 244, 367, 390 
in cartesian coordinates, 370 
in curvilinear coordinates, 407 
Graeffe’s root-squaring method, 679n. 
Gram-Schmidt method, 351 
Graphical solution of equations, 678 
Gravitational attraction, 277p., 409 
motion under, 47, 49p. 
Gravitational constant, 46 
Gravitational field, 407 
Gravitational potential, 409 
Gravity, center of, 275, 281 
Gravity dam, 593 
Greatest lower bound, 773 
Green’s function, 501 
for half space, 502 
Green’s identities, 391p., 493 


Green’s theorem, in plane, 391, 402p. 

symmetric forms of, 391p., 493 
Growth factor, 418 


Harmonic analysis, 711 
Harmonic function, 468, 560, 585 
average value theorem for, 498 
conjugate, 560, 588 
differentiability of, 559 
maximum values of, 499p., 506, 558, 561 
Harmonics, 177 
Heat capacity, 455 
Heat equation, 414, 455 
solution of, by integrals, 482 
by separation of variables, 459 
by series, 455-471 
uniqueness of, 466, 506 
Heat flow, 414, 455-467, 483, 504, 512 
connection with random walks, 653 
in a rod, 456-466, 489, 769 
source function for, 491, 653 
in a sphere, 471 
Heat source, 489, 504, 653 
Heaviside’s expansion theorem, 766 
Helix, 314, 316p. 

Helmholtz formula, 499 
Hermitian form, 348 
Hermitian matrix, 349 
Holomorphic function, 543 
Homogeneous differential equations (see 
Ordinary differential equations) 
Homogeneous functions, 18, 234 
Euler’s theorem on, 234 
Hooke’s law, 80 
Horner’s method, 679n. 

Hydrodynamics, 416, 419 
(See also Fluid flow) 

Hydrostatic pressure, 593 
Hyperbolic differential equation, 507 
difference equation for, 513 
Hyperbolic functions, 537, 591 
Hypergeometric equation, 165 

Ideal fluid, 419 

Images, method of, 448, 462p. 

Implicit functions, differentiation of, 280, 
235 

Improper integrals (see Integrals) 

Impulse function, 759 
Indefinite integral, 551 



INDEX 


805 


Independence, linear, 52, 70, 317 
of path, 378, 393 

Independent events in probability, 610, 
618, 638 

Indidal equation, 161 
Inertia, moment of, 274 
Infinite series (see Series) 

Inner product (see Scalar product) 
Integral calculus, fundamental theorem of, 
9, 261, 550 
Integral curve, 7 
Integral equations, 767 
Integrals, of analytic functions, 545, 547, 
551 

of Cauchy’s type, 557 
change of variables in, 270 
of complex functions, 545 
contour (see Line integrals) 
convergence of, absolute, 755 
differentiation of, 261, 262 
elliptic, 49, 86p., 148 
evaluation of, by fundamental theorem, 
9, 261 

by numerical methods, 717 
by residue theorem, 599 
by series, 147 
improper, 118, 553n., 602 
principal value of, 602 
indefinite,^ 551 
Lebesgue, 774 
line (see Line integrals) 
mean-value theorem for, 380 
multiple, 270 

particular (see Particular integrals) 
probability, 776 
Riemann, 771 
Stieltjes, 630 
surface, 277, 373 
transformation of, 382-402 
volume, 374 
Integrating factor, 22 
Integration, numerical, 715 
Interpolation, 679 
Interpolation formulas, 696, 699 
Interval, closed, 132n., 772n. 
of convergence, 139 
open, 218, 772n. 

Inverse elementary functions, 539p., 540p. 
Inversions, of matrices, 333, 350 
of order, 742 
Irrotationa] field, 402 
Irrotational flow, 412 


Isoclines, 36 

Isolated singular points, 569 

Iterative methods, 679, 684, 689, 721-730 

( see Bessel’s functions) 

Jacobian, 238, 242p., 271 
Jump of a function, 520 

K»(x) (see Bessel’s functions) 

Kinetic energy, 43, 306p. 

Kronecker delta, 321 

Lagrange’s differential equation, 26 
Lagrange’s interpolation formula, 699 
Lagrange's multipliers, 250, 254 
Laplace transform, 754r~769 
bilateral, 762n. 
convolution theorem for, 762 
of derivatives, 756 
of Dirac’s “function/' 759 
Heaviside’s theorem on, 766 
solution by, of differential equations, 
756-762 

of integral equations, 767 
tables of, 770 
unilateral, 762n. 

Laplace’s difference equation, 511, 735 
Laplace’s equation, 409, 413, 416, 419, 464, 
467, 735 

Laplace's law in probability, 647 
Laplace-de Moivre limit theorem, 648 
Laplacian operator, 387 
in curvilinear coordinates, 407 
Laurent’s expansion, 564 
uniqueness of, 567 
Laurent’s theorem, 565 
Law, of errors, 662 
of large numbers, 650 
of mechanics, 302 
Newton’s (see Newton's law) 
parallelogram, 288, 529 
of probability, binomial, 639, 647 
normal, 647, 653 
of reflection, 297 p. 
of refraction, 297p. 
of small numbers, 654 
Least squares, 663p., 702 
connection with orthogonal functions, 
200 

curve fitting by, 702 



806 


INDEX 


Lebeegue integral, 774 
Lebeague theorem, 775 
Legendre polynomials, 150 
expansion in series of, 196, 473 
generating function for, 159p. 
orthogonality of, 198 
Rodrigues’ formula for, 159p. 
Legendre's equation, 158 
Leibniz' formula, for differentiation of 
integrals, 262 

use in evaluating integrals, 263 
Leibniz’ test, 128 
Length of arc, 301, 362 
of an ellipse, 147 
Level surface, 369 
Line, equation of, 306 
Line integrals, 373 
of analytic functions, 547-554 
in complex plane, 545 
independent of path, 378, 393 
transformation of, 382-402 
of vector functions, 374 
Linear algebraic equations, 350, 687, 689, 
749 

Linear dependence, 52, 70 
of vectors, 317 

Linear differential equations (see Ordinary 
differential equations) 

Linear fractional transformation (bilinear 
transformation), 577 p. 

Linear operators, 336, 754 
linear transformation, 332 
linear vector spaces, 316 
Linearity, property of, 51 
Lipschitz condition, 38 
Log z, 537, 681 

Logarithmic function, 537, 581 
principal value of, 537 
Lower bound, 773 


M test, 134 

Maclaurin’s formula, 260 
Maclaurin's series, 144 
Mapping, by analytic functions, 575-594 
conformal (see Conformal mapping) 
Mass, center of, 303 
motion of, 44, 304 

Matrices, algebraic operations on, 327-331 
inversion of, 333, 350 
product of, 328 
transformation of, 340*350 


Matrix, 327, 749 
augmented, 750 
characteristic equation of, 344 
characteristic values of, 344 
conjugate, 343 
determinant, of, 750 
diagonal, 329, 339, 343 
Hermitian, 349 
identity, 329 
inverse of, 333 
orthogonal, 340 
rank of, 330, 750 
scalar, 329 
singular, 331 
square, 327 
symmetric, 347 
transpose of, 334 
unit, 329 

unitary, 343 p., 350 
zero, 329 

Maxima and minima, 246 
absolute, 247 
constrained, 249, 269 
relative, 247 

(See also Calculus of variations) 
Maximum modulus theorem, 558 
Maximum principles, 506, 507 
Maxwell-Boltzmann distribution, 633 
Mean errors, 200, 659, 660 
reliability of estimate of, 670 
Mean-value theorem, of differential cal- 
culus, 224 
for integrals, 380 
Measurable set, 773 
Measure, 772 

Measure numbers, 291, 319 
Measure theory, 614, 772 
Mechanics, laws of, 302 
Median value, 637p. 

Membrane, under gas pressure, 482p. 
vibration, of circular, 480 
of rectangular, 474 
Metric coefficients, 360 
Minima (see Maxima and minima) 
Minimax, 249 
Minimizing curve, 265 
Minimum descent, curve of, 269p., 767 
Minimum potential-energy principle, 264 
Minors of a determinant, 741 
Modes, 477, 481 

Modulus of a complex number, 528 
Moment, bending, 4 35p. 



mass 


807 


Moment, of dipole, 497 
of force, 303 
of inertia, 274 
of momentum, 305 
Momentum, angular, 305 
linear, 42, 44 
moment of, 305 
Momentum vector, 305 
Monte Carlo methods, 652 
Multiple integrals, 270 
Multiply connected region, 383 
Mutually exclusive events, 610, 619 


Nabla or del, V, 370 
Neighborhood of a point, 540 
Neumann’s function, 503 
for half plane, 504 
Neumann’s problem, 503 
Newtonian potential, 277 p. t 409 
Newton’s interpolation formulas, 696 
Newton's law, of attraction, 46, 409 
of cooling, 461 
of gravitation, 46, 408 
of motion, 42, 43 

Newton's method of solving equations, 684 
Nodal lines, 477 
Nodes, 429, 445 
Normal, to a curve, 311 
principal, 311 
to a surface, 309, 369 
Normal acceleration, 313 
Normal derivatives, 244, 369 
Normal distribution, 633, 651 
bivariate, 666 
Normal equations, 703 
Normal law of probability, 647 
interpretation of, 653 
Normal line, 309 
(See also Normal) 

Numerical analysis, 673-736 
Numerical differentiation, 698 
Numerical integration, 715 
Numerical solution of differential equa- 
tions, 37, 721-736 


Odd functions, 183 
Fourier expansion for, 185 
Operator, curl, 398, 407 
A 57, 430p. 

V, 370, 386, 407, 692 


Operator, V 2 , 387 
A, 510, 692 
difference, 510, 692 
div, 386, 406 
Fourier transform, 482 
Laplace, 387 
Laplace transform, 754 
linear, 336, 754 

Order, of differential equations, 6, 29, 76, 
425 

interchange in partial differentiation, 
221 

inversions of, 742 
reduction of, 29, 76 
Ordinary differential equations, 1-106 
Abel's theorem for, 54 p. 

Bernoulli’s, 27 
Bessel's, 159 

boundary-value problems in, 91, 730 
Cauchy’s, 78 

Chaplygin's method for, 37 
characteristic equation for, 54, 67, 101 
Clairaut’s, 27p. 

with constant coefficients, 54, 66, 100 
of electric circuits, 81, 100, 761p. 
Euler’s, 267 
Euler-Cauchy’s, 78 
exact, 20 

existence and uniqueness theorems for, 
5, 7, 38, 157 
first-order, 17-50 
linear, 23, 51, 59 
Fuchs’ theorem on, 157 
Gauss’ hypergeometric, 165p. 
homogeneous, first-order, 18 
linear, 51, 54, 59, 96 
systems of, 100, 733 
hypergeometric, 165 
indici&l equations for, 161 
initial-value problem for, 9, 90, 730 
integral curves for, 7 
integrals of, 7 
integrating factors for, 22 
integration between limits, 13 
isoclines for, 36 
Lagrange’s, 26 
Legendre’s, 158 

linear, complementary function for, 59a. 
with constant coefficients, 54, 66 
systems of, 95, 733 

with variable coefficients, 51, 59, 
70, 153 



INDEX 


808 

Ordinary differential equations, order of, 
6, 425 

reduction of, 29, 76 
with separable variables, 18 
Solutions of, 6 
general, 10, 52, 59, 67, 102 
by Laplace transform, 756-768 
linearly independent, 7, 59, 72, 76 p. 
by numerical methods, 87, 721-736 
particular, 7, 59, 72, 76 p. 
by power series, 153 
singular, 8, 34 
stability of, 103, 105 
uniqueness of, 7, 38, 157 
systems of, 95, 110, 727 
characteristic equation for, 101, 733 
Origin, 288 

Orthogonal coordinates, 363 
Orthogonal curves, 31 
Orthogonal matrices, 340 
Orthogonal sets of functions, 195 
completeness and closure of, 203 
expansion in series of, 201 
relation of least squares to, 202 
Orthogonal trajectories, 30 
Orthogonal transformations, 340 
Orthogonal vectors, 319 
Orthogonality weighted, 197 
Orthogonalization, of matrices, 340, 350 
of vectors, 320 

Orthonormal functions, 195, 197 
Orthonormal vectors, 320 
Oscillations, of cable, 445, 454 
damped, 449 
period of, 44, 81, 84 
of spring, 80, 82, 86, 88, 89 
Osculating plane, 311 


Parabolic coordinates, 408p. 

Parabolic differential equation, 506 
difference equation for, 512 
Parabolic mirror, 33 
Parallelogram law of addition, 288, 529 
Parsev&l's equality, 202, 204p. 

Partial differential equations, 5, 425-521 
boundary conditions for, 443 
canonical forms of, 517 
characteristic values for, 507p. 
characteristics for, 441, 521 
of elliptic type, 504 
of heat flow 414, 455 


Partial differential equations, of hydrody- 
namics, 416, 429 
of hyperbolic type, 504, 516 
of parabolic type, 504, 516 
of potential theory, 409, 41 In. 
solutions of, by Fourier transform, 
482-490 

fundamental, 521 
by integrals, 482-504 
by Laplace transform, 769 p. 
numerical, 734 
by series, 448-482 
uniqueness of, 505-510 
of vibrating membranes, 475, 480 
of vibrating rods, 435p., 485p. 
of vibrating string, 431, 484 
of wave motion, 428 
Partial differentiation, 219 
interchange of order in, 221 
Partial sum of series, 111 
Particular integrals, 7 
by method of undetermined coefficients, 
59 

by variation of parameters, 72 
Pendulum, 48, 49, 85 p. 

Period, of oscillations, 44, 81, 84 
of pendulum, 49, 85 p. 
of vibration, 443 
of waves, 428 
Permutations, 611 

Phase of simple harmonic motion, 44 
Picard’s method, 509 
Piecewise continuity, 755 
Piecewise smoothness, 206, 372 
Plancherel’s theorem, 483n. 

Plane, equation of, 306 
osculating, 311 
tangent, 277 n., 309 
Point at infinity, 577 
Point set, 773 
measure of, 774 
Point vortex, 591 

Points, in n-dimensional space, 316 
in sample space, 613 
Poisson's distribution, 633 
Poisson's equation, 268, 41 In., 495 
uniqueness of solution of, 499p. 

Poisson's formula, 495 
for a circle, 470, 599 p. 
for a half plane, 490 
for a half space, 503 
Poisson's law of probability, 654 



INDEX 


809 


Poles of analytic functions, 570, 574 
residues at, 571 
simple, 570 

Polynomial representation of data, 694, 
702 

Potential, complex, 587 
electrostatic, 467, 496 
gravitational, 409, 468 
Newtonian, 277 p M 409 
Potential energy, 43 
principle of minimum, 264 
Potential theory, 468 

(, See also Dirichlet's problem; Neu- 
mann's problem) 

Power series, 138 
Abel’s theorem on, 142 

convergence of, absolute and uniform, 
139 

radius of, 142, 170, 562 
differentiation and integration of, 140 
evaluation of integrals by, 147 
expansions m, 144, 561 
multiplication of, 142 
solution of differential equations by, 153 
substitution in, 152 

uniqueness of representation by, 141, 
146 

Precision constant, 662, 670 
Pressure, atmospheric, 50 
on gravity dam, 593 
in star's interior, 48 
Primitive, 551 
Principal argument, 528 
Principal normal, 31 1 
Principal value, of improper integrals, 602 
of log z, 537 

Probability, 610, 614, 652 
binomial law of, 639, 647 
compound and total, theorems on, 617, 
635 

events in, 610, 618, 619, 638 
Laplace's law in, 647 
law of large numbers, 651 
law of small numbers (Poisson's), 654 
marginal, 625 
normal law of, 647, 653 
Probability density, 632, 634 
Gauss', 633 
joint, 634 

Maxwell-Boltzmann’s, 633 
Poisson's, 633 

Probability integral, table of, 776 


Probable error, 662 
Probable value, 635 
Product, of determinants, 325, 745 
of matrices, 328 
of vectors, 293, 295, 298 
Projectiles, 50 

Pulley, slipping of belt on, 1 1 
Pursuit curves, 33 
Pythagorean formula, 322 


Quadratic forms, 347 
differential, 362 
positive definite, 349 
Quadrature, 715 


Radiation from antenna, 486 
Radiation condition, 501 
Radius of convergence, 139, 170, 562 
Random molecular motions, 653n. 
Random process, 622 
Random variables, 623, 662 
Random walks, 653 
Rank of a matrix, 330, 750 
Ratio test, 125 
Rational number, 772 
Reflection, law of, 297 p, 
transformation of, 341 
Refraction, law of, 297p. 

Regions, bounded, 535 
closed, 218, 535 
connected, 383 
multiply, 383, 535 
simply, 383, 535 
finite, 535 
open, 218 
regular, 383 

Relative frequency, 615, 638, 642 
Remainder in Taylor's series, 144 
Residuals, 703 
Residue theorem, 573 
evaluation of real integrals by, 599 
Residues, 571 
Resonance, 86, 89, 453 
Resonant frequency, 89, 477 
Riemann function, 508 
Riemann integral, 771 
Riemann 's mapping theorem, 586 
Rocket, motion of, 45, 46 p. 
thrust, 45 

Rodrigues’ formula, 159p. 



810 


INDEX 


Roots of unity, 532 

Rotation, of shaft, critical speed of, 94 
transformation of, 341 
velocity of, 399p. 


Sample space, 613, 634 
Scalar, 287 
Scalar fields, 367 
Scalar product, 293n., 319 
Scalar triple product, 297 
Schwarz’ inequality, 322 
Schwarz-Christ off el mapping formula, 599 
Seidel’s method, 689 
Separation of variables, 18, 459, 732 
Series, 111 
addition of, 117 
alternating, 128 
basic properties of, 116 
binomial, 155 
Cauchy’s criterion for, 115 
comparison tests for, 122, 125, 134 
of complex terras, 169 
convergence of, 112 
absolute, 127, 170 
conditional, 129 
fundamental principle for, 114 
in the mean, 200 
pointwise, 204 

tests for, comparison, 113, 122 
integral, 118 
Leibniz', 128 
ratio, 125 
uniform, 132, 136 
Weierstrass test for, 134 
differentiation of, 135 
evaluation of integrals by, 147 
Fourier (see Fourier series) 
geometric, 115 
harmonic, 113 
integration of, 135 
Laurent's, 565 
Maclaurin’s, 144 
multiplication of, 131 
of orthogonal functions, 195 
power (see Power series) 
rearrangement of, 129 
remainder in, 113 

solution of differential equations by, 
153, 465 
sum of, 112 
partial, 111 


Series, Taylor’s, 144, 561 
telescoping, 116p. 
trigonometric (see Fourier aeries) 

Shaft, critical speed of rotation, 94 
Shearing load, 435p. 

Significance level, 649 
Similar transformations, 339 
Similitude, principle of, 433 
Simple closed curve, 383n. 

Simple harmonic motion, 44 
Simple pendulum, 48 
Simply connected region, 383 
Simpson’s rule, 717 
Singular integral, 493 
Singular points, 543, 569 
essential, 570, 574 
isolated, 569 

Singular solutions of differential equa 
tions, 34 
Sink, 384, 591 
Solenoidal field, 402 
Solid angle, 399p. 

Solution of differential equations («e< 
Ordinary differential equations; Par 
tial differential equations) 

Sommerf eld’s radiation condition, 501 
Sound, equation of propagation of, 420 
velocity of, 435p. 

Source, 384, 591 
of heat, 489, 504, 653 
Space, complex, 322 
dimensionality of, 317 
Euclidean, 321, 374n. 
linear vector, 316 
sample, 613, 634 

Space curves (Frenet-Serret formulas), 315 
Specific heat, 415 
Spectral theory, 199 
Spherical coordinates, 360, 367p. 

Spring, oscillation of, 80, 82, 86, 88, 99 
Stability, of columns, 91 
of rotating shafts, 92 
of solutions of differential equations, 
103, 105 

Stagnation points, 590 
Standard deviation, 664, 669 
of the mean, 667 
unbiased estimate of, 670 
Statistical hypothesis, 614 
relation to significance level, 649 
Steady-state solutions, 88, 765 
Steady-state temperature, 457, 471 



INDEX 


811 


Stieitjaa integral, 630 
Stirling’s formula, 644 
Stokes* theorem, 400, 40 2p. 

Stream function, 413 
Streamlines, 413, 414p,, 587 
String, vibration of, 425, 431-454, 484 
Sturm-Liouville theory, 199 
Summation convention, 324 
Superposition principle, 766 
Surface, level, 369 
normal to, 309, 369, 373n. 
one-sided, 383 
piecewise smooth, 277, 383 
regular, 384 

tangent plane to, 277 n., 309 
two-sided, 277, 373n., 383 
Surface integrals, 277, 373 
Systems of equations, differential, 95, 733 
linear algebraic, 350, 687, 689, 749 


Tangent line, 301, 31 1 
Tangent plane, 277 n., 309, 311 
Tautochrone, 767 
Taylor's formula, 143 
approximations by. 149 
for functions of several variables, 257 
Taylor’s series, 144, 561 
Telegrapher’s equation, 514 
Thermal conductivity, 415, 455 
Torque, 303 

(See also Moment) 

Torsion, 312 

Total differential, 226, 234 
Total probability, 618. 635 
Transcendental equations, 677, 679. 684 
Transforms, Fourier, 194 
Laplace, 754 
Trapezoidal rule, 717 
Trigonometric functions, 172, 536 


Undetermined coefficients, method of, 59 
Uniform convergence, 132 
Uniqueness, of representation, in Fourier 
series, 186 

in Laurent's series, 567 
in power series, 141, 146 
of solutions, of ordinary differential 
equations, 7, 38, 157 
of partial differential equations, 505- 
510 


Unit impulse, 759 
Unit vectors (see Base vectors) 
Unitary matrices, 343p., 350 
Unitary transformations, 343p. 


Value, absolute, of complex numbers, 168, 
528 

characteristic, 344, 507p., 732 
expected, 623, 639 
extreme, 250, 264 

maximum, of harmonic function, 499p , 
506, 558, 561 
median, 637p. 

principal, of improper integrals, 602 
of logarithmic function, 537 
probable, 635 
Variance, 664, 669 
Variation, 265 
of parameters, 72 
Vector, velocity. 301, 367p. 

Vector acceleration, 302p., 313, 307p. 
Vector analysis, 285-361 
( See also Vector field theory) 

Vector field theory, 355-420 
Vector functions, continuity of. 299, 368 
line integrals of, 374 
Vector product, 295 
Vector spaces, 316, 323 
Vectors, algebraic operations on, 288-298, 
316-324 

base or coordinate (see Base oi coordi- 
nate vectors) 
bound, 288 
characteristic, 344 
components of 291, 317, 321 
continued products of, 297 
coordinate, 319 
differentiation of, 299 
free, 288 

linear dependence of, 317 
magnitude of, 288 
momentum, 305 
orthogonal, 319 
orthonormal, 320 
parallelogram law for, 288 
product of, 293, 295, 298 
sliding, 288 
unit, 319 
zero, 289, 317 
Velocity, angular, 302 
of escape, 47 



INDEX 


8X2 

Velocity, of rotation, 399 p. 
of sound, 435p. 
in wave motion, 434 
Velocity potential, 413, 419, 584, 587 
Velocity vector, 301, 367p, 

Vibration, of beams, 435 p. 
of membranes, 474, 480 
period of, 443 
of string, 425, 431-454, 484 
Viscosity of gases, 450 
Viscous damping, 82, 80, 87, 449 
Volume integral, 270, 374 
Vortex, 591 


Wave equation, 428, 432, 484, 499, 508 
with damping, 449, 452 
solution of v D'Alembert's, 439, 485 
Fourier integral, 484 


Wave equation, solution of, Fourier series, 
448 

by separation of variables, 463p. 
uniqueness of, 439, 442, 508 
Wave front, 486, 519 
Waves, 425, 436 
amplitude of, 428 
period of, 428 
plane, 488 
shock, 520 
standing, 428 
Weierstrass M test, 134 
Work, 302, 306p., 409, 418 
Wronskian determinant, 52, 54p., 71 


Zero vector, 289, 317 
Zonal harmonics (see Legendre poly- 
nomials) 






DELHI POLYTECHNIC 

LIBRARY 

* Cl ASS NO. $ f I 

BOOK NO. H 4-7 R 
ACCESSION NO. A ' €